You are on page 1of 94

TM 6004 – Teknik Pemboran Lanjut

RESUME BAB III BUKU ZOBACK


BASIC CONSTITUTIVE LAWS

Babas Samudera Hafwandi (22220003)


Program Studi Magister Teknik Perminyakan ITB
1
OUTLINE
1. Introduction
2. Linear Elasticity
3. Elastic Moduli and seismic wave velocity
4. Elasticity anisotropy

2
OUTLINE (Continued)
5. Poroelasticity and Effective Stress
6. Poroelasticity and Dispersion
7. Viscous Deformation in Uncemented Sands
8. Thermoplasticity
9. Conclusions.

3
1. Introduction
• There are 4 generic types of constitutive laws for homogeneous
and isotropic materials.
1. Elastic constitutive laws
2. Poroelastic constitutive laws
3. Elastic-plastic constitutive laws
4. Viscoelastic constitutive laws.

4
1.1. Elastic Constitutive Laws
• A linearly elastic material (Figure 1) is one in which stress and
strain are linearly proportional and deformation is reversible.
• This can be conceptualized in terms of a force applied to a
spring where the constant of proportionality is the spring
constant, k.

5
1.1. Elastic Constitutive Laws (Continued)
• An ideal elastic rock strains linearly in response to an applied
stress in which the stiffness of the rock is E, Young’s modulus.
• An actual rock mechanics test is presented in Figure 2. to
illustrate how a relatively well-cemented sandstone exhibits
nearly ideal elastic behavior over a considerable range of
applied stresses.

6
1.1. Elastic Constitutive Laws (Continued)

Fig 1. Schematic illustration of elastic constitutive laws (Zoback.,


2007)
7
1.1. Elastic Constitutive
Laws (Continued)
Stress, MPa

Fig 2. Typical laboratory


stress–strain data for a well-
cemented rock being deformed
uniaxially (Zoback., 2007)
Strain 8
1.2. Poroelastic Constitutive Laws
• A porous rock saturated with fluid will exhibit poroelastic
behavior.
• One manifestation of poreoelasticity is that the stiffness of a
fluid-saturated rock will depend on the rate at which external
force is applied.
• When force is applied quickly, the pore pressure in the rock’s
pores increases because the pore fluid is carrying some of the
applied stress and the rock behaves in an undrained manner.

9
1.2. Poroelastic Constitutive Laws
(Continued)
• In other words, if stress is applied faster than fluid pressure can
drain away, the fluid carries some of the applied stress and the
rock is relatively stiff.
• However, when an external force is applied slowly, any increase
in fluid pressure associated with compression of the pores has
time to drain away such that the rock’s stiffness is the same as if
no fluid was present.

10
1.2. Poroelastic Constitutive Laws
(Continued)

Fig 3. Schematic illustration of poroelastic constitutive laws


(Zoback., 2007)
11
1.3. Elastic-Plastic Constitutive Laws
• Figure 4. illustrates elastic–plastic behavior. In this case, the
rock behaves elastically to the stress level at which it yields and
then deforms plastically without limit.
• Upon unloading the rock would again behave elastically.
Although some highly deformable rocks behave this way in
laboratory testing (right panel).
• In this case, appreciable deformation (i.e. fault slip) can occur at
a relatively constant stress level (i.e. that required to cause
optimally oriented faults to slip).
12
1.3. Elastic-Plastic Constitutive Laws
(Continued)

Fig 4. Schematic illustration of elastic-plastic constitutive laws


(Zoback., 2007)
13
1.4. Viscoelastic Constitutive Laws
• A viscoelastic rock (Figure 5) is one in which the deformation in
response to an applied stress or strain is rate dependent.
• The stress required to cause a certain amount of deformation in
the rock depends on the apparent viscosity, η, of the rock (center
panel).
• One can also consider the stress resulting from an
instantaneously applied deformation (right panel) which will
decay at a rate depending on the rock’s viscosity.

14
1.4. Viscoelastic Constitutive Laws
• The conceptual model shown in the left panel of Figure 5.
corresponds to a specific type of viscoelastic material known as
a standard linear solid.
• A viscous material that exhibits permanent deformation after
application of a load is described as viscoplastic.

15
1.4. Viscoelastic Constitutive Laws
(Continued)

Fig 5. Schematic illustration of viscoelastic constitutive laws


(Zoback., 2007)
16
2. Linear Elasticity
• The theory of elasticity typically in terms of infinitesimally
small deformations.
• In this case, no significant damage or alteration of the rock
results from an applied stress and the assumption that stress and
strain are linearly proportional and fully reversible is likely to be
valid.

17
2. Linear Elasticity (Continued)
In such a material, stress can be expressed in terms of strain by
the following relation:

.............(1)

Where:

18
2. Linear Elasticity (Continued)
Sij = Stress tensor, psi
λ = Lame constant, psi
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j)
ε00 = Volumetric strain, volume
G = Shear modulus, psi
εij = Strain tensor, volume.

19
 𝛿 𝑢 1
Axial Strain Ɛ11 𝛿 𝑥1
 𝛿 𝑢 3
2. Linear Elasticity
Lateral Expansion Ɛ33 𝛿 𝑥3 (Continued)
 X 1
𝛿  𝑢 1

 𝛿 𝑢 1  𝛿 𝑢 3
3 2 𝛿  x 1

 X 3
𝛿  𝑥 3
Young Modulus (E) Poisson Ratio (v) Figure 6. Concept of Young
  𝑆 11  
v Modulus and Poisson Ratio
𝐸=
Ɛ 11 (Zoback, 2007)
20
 
Ɛ 31=
𝛿 𝑢3 2. Linear Elasticity
Shear Strain 𝛿 𝑢1 (Continued)
 X 1 𝛿  𝑢 3

𝛿  x 1

 X 3

Shear Modulus (G)


Figure 7. Concept of Shear
  1 𝑆 13
G= ( ) Modulus (Zoback, 2007)
2 Ɛ 13 21
Volumetric Strain
Ɛ00 = Ɛ11 + Ɛ22 + Ɛ33
2. Linear Elasticity
 X 1 (Continued)
 𝛿 u 1

 X 3

𝛿  x 2  𝛿 𝑢 3
 X 2
Bulk Modulus (K)
Figure 8. Concept of Bulk
  𝑆 00
K =( ) Modulus (Zoback, 2007)
Ɛ 00 22
2. Linear Elasticity (Continued)
• An important aspect of the theory of elasticity in homogeneous,
isotropic material is that only two elastic moduli are needed to
describe fully material behavior.
• Strain can be expressed in terms of applied stress utilizing the
following relation :

............(2)

23
2. Linear Elasticity (Continued)
• Because only two elastic moduli are needed, it is often
convenient to express elastic moduli with respect to each other.
• For example, if one has knowledge of the bulk modulus and
Young’s modulus, one can compute the shear modulus from :

............(3)

24
2. Linear Elasticity (Continued)
Sij = Stress tensor, psi
λ = Lame constant, psi
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j)
εij = Strain tensor, volume
G = Shear modulus, psi
S00 = Volumetric strain, volume

25
2. Linear Elasticity (Continued)
K = Bulk modulus, volume
E = Young’s modulus, psi.

26
3. Elastic Moduli and Seismic Wave
Velocity
• In an elastic, isotropic, homogeneous solid the elastic moduli
also can be determined from the velocity of compressional
waves (Vp) and shear waves (Vs) using the following relations:

………….. (4)

27
3. Elastic Moduli and Seismic Wave
Velocity (Continued)
Where:
Vp = Compressional waves, ft/s
Vs = Shear waves, ft/s
K = Bulk modulus, volume
ρ = Fluid density, lbm/ft3
G = Shear modulus, psi.

28
3. Elastic Moduli and Seismic Wave
Velocity (Continued)
• It is also sometimes useful to consider relative rock stiffnesses
directly as determined from seismic wave velocities. For this
reason the so-called M modulus has been defined:

....................(5)

29
3. Elastic Moduli and Seismic Wave
Velocity (Continued)
Where:
M = M modulus, psi
Vp = Compressional waves, ft/s
K = Bulk modulus, volume
ρ = Fluid density, lbm/ft3
G = Shear modulus, psi.

30
3. Elastic Moduli and Seismic Wave
Velocity (Continued)
• Poisson’s ratio can be determined from Vp and Vs utilizing the
following relation:

....................(6)

31
3. Elastic Moduli and Seismic Wave
Velocity (Continued)
Where:
ν = Poisson’s ratio, dimensionless
Vp = Compressional waves, ft/s
Vs = Shear waves, ft/s.

32
4. Elasticity Anisotropy
• A number of factors can make a rock mass anisotropic :
1.Aligned microcracks
2.High differential stress
3.Aligned minerals (such as mica and clay) along bedding
planes
4.Macroscopic fractures and faults.

33
4. Elasticity Anisotropy (Continued)
• Elastic anisotropy can have considerable effects on seismic
wave velocities, and is especially important with respect to
shear wave propagation.
• With respect to elastic anisotropy, the general formulation that
relates stress to strain is:

.........(7)

34
2. Linear Elasticity (Continued)
Sij = Stress tensor, psi
cijkl = Elastic stiffness matrix, lb/inch
εkl = Matrix strain, volume .

35
5. Poroelasticity and Effective Stress
• In a porous elastic solid saturated with a fluid, the theory of
poroelasticity describes the constitutive behavior of rock.
• The three principal assumptions associated with this theory are
similar to those used for defining pore pressure :

36
5. Poroelasticity and Effective Stress
(Continued)
1.First, there is an interconnected pore system uniformly
saturated with fluid.
2.Second, the total volume of the pore system is small compared
to the volume of the rock as a whole.
3.Third, consider pressure in the pores, the total stress acting on
the rock externally and the stresses acting on individual grains
in terms of statistically averaged uniform values.

37
5. Poroelasticity and Effective Stress
(Continued)
• The concept of effective stress is based on the pioneering work
in soil mechanics by Terzaghi (1923) who noted that the
behavior of a soil (or a saturated rock) will be controlled by the
effective stresses, the differences between externally applied
stresses and internal pore pressure.
• The so-called “simple” or Terzaghi definition of effective stress
is :

……..(8)
38
5. Poroelasticity and Effective Stress
(Continued)
σij = Effective stress, psi
Sij = Stress tensor, psi
Pp = Pore pressure, psi
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j).

39
5. Poroelasticity and Effective Stress
(Continued)
• Empirical data have shown that the effective stress law is a
useful approximation which works well for a number of rock
properties, but for other rock properties, the law needs
modification.
• Nur and Byerlee (1971) proposed an “exact” effective stress
law, which works well for volumetric strain. In their formulation
:

40
5. Poroelasticity and Effective Stress
(Continued)
........(9)

Where α is the biot parameters :

.............(10)

Where:

41
5. Poroelasticity and Effective Stress
(Continued)
σij = Effective stress, psi
Sij = Stress tensor, psi
Pp = Pore pressure, psi
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j).
Kb = Drained bulk modulus of the rock or aggregate
Kg = Bulk modulus of the rock’s individual solid grains.

42
5. Poroelasticity and Effective Stress
(Continued)
• Thus, to consider the effect of pore fluids on stress we can re-
write equation (1) as follows:

……….(11)

such that the last term incorporates pore pressure effects.

43
5. Poroelasticity and Effective Stress
(Continued)
Sij = Stress tensor, psi
λ = Lame constant, psi
ε00 = Volumetric strain, volume
P0 = Pore pressure, psi
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j).
εij = Strain tensor, volume
G = Shear modulus, psi.

44
6. Poroelasticity and Dispersion
• The stiffness (elastic moduli) of a poroelastic rock is rate
dependent. In regard to seismic wave propagation, this means
that P-wave and S-wave velocities will be frequency dependent.
• Figure 9a illustrates the difference between laboratory bulk
modulus measurements of an uncemented Gulf of Mexico sand
determined statically, and using ultrasonic (∼1 MHz) laboratory
velocity measurements.

45
6. Poroelasticity and Dispersion
(Continued)
• Note that at low confining pressure, there is about a factor of 2
difference between the moduli determined the two different
ways.
• As confining pressure increases the difference increases
significantly. Thus, there can be significant differences in
velocity (or the elastic modulus) depending on the frequency of
seismic waves.

46
6. Poroelasticity and Dispersion
(Continued)
• Seismic-wave frequencies typical of a reflection seismic
measurement (∼10–50 Hz) are slower (yield lower moduli) than
sonic logs (typically∼10 kHz), and sonic logs yield slower
velocities than ultrasonic laboratory measurements (typically
∼1 MHz).
• As illustrated in Figure 9b, this effect is much more significant
for P-wave velocity than S-wave velocity.

47
6. Poroelasticity and Dispersion (Continued)

Figure 9a. Comparison between


static bulk modulus measurements
and dynamic (ultrasonic)
measurements for a dry,
uncemented sand as a function of
pressure (Zoback., 2007).
48
6. Poroelasticity and Dispersion (Continued)

Figure 9b. Comparison between P


and S-wave velocity measurements
in a water saturated at frequencies
corresponding to geophysical logs
(several kilohertz) and laboratory
measurements (∼one megahertz)
(Zoback., 2007).
49
6. Poroelasticity and Dispersion
(Continued)
• There are a number of different processes affecting seismic
wave propagation that contribute to the effects shown in Figure
9:

1. The seismic waves associated with seismic reflection


profiling, well logging and laboratory studies sample very
different volumes of rock.

50
6. Poroelasticity and Dispersion
(Continued)
2. When comparing static measurements with ultrasonic
measurements, it is important to remember that the amount
of strain to which the samples are subjected is markedly
different, which can affect the measurement stiffness
3. Pore fluid effects can contribute dramatically to dispersion
at high frequencies.

51
6. Poroelasticity and Dispersion
(Continued)
• SQRT (squirt, or local flow) is a theory used to explain the
dependence of wave velocity on frequency in a saturated
poroelastic rock at high frequency.
• Fundamentally, SQRT (and theories like it) calculates the
increase in rock stiffness (hence the increase in velocity)
associated with the amount of pressure “carried” by pore fluids
as seismic waves pass through rock.
• At very high (ultrasonic) frequencies, there is insufficient time
for localized fluid flow to dissipate local pressure increases.

52
6. Poroelasticity and Dispersion
(Continued)
• Hence, the rock appears quite stiff (corresponding to the
undrained modulus and fast ultrasonic P-wave velocities as
measured in the lab) because the pore fluid pressure is
contributing to the stiffness of the rock.
• Conversely, at relatively low (seismic or well logging)
frequencies, the rock deforms with a “drained” modulus.

53
6. Poroelasticity and Dispersion
(Continued)
• Hence, the rock is relatively compliant (relatively slow P-wave
velocities would be measured in situ).
• It is intuitively clear why the permeability of the rock and the
viscosity of the fluid affect the transition frequency from
drained to undrained behavior.

54
6. Poroelasticity and Dispersion
(Continued)
• This is illustrated in Figure 9b. SQRT theory predicts the
observed dispersion for a viscosity of 1 cp, which is
appropriate for the water filling the pores of this rock.
• Had there been a more viscous fluid in the pores (or if the
permeability of the rock was lower), the transition
frequency would shift to lower frequencies, potentially
affecting velocities measured with sonic logging tools
(∼104 Hz).

55
6. Poroelasticity and Dispersion
(Continued)
• Figure 10a. clearly demonstrates the difference between static
and dynamic bulk modulus in an uncemented Gulf of Mexico
sand.
• As shown by the hydrostatic loading cycles, the static stiffness
(corresponding to the slope of the loading line) is much lower
than the dynamic stiffness (indicated by the slope of the short
lines) determined from ultrasonic velocity measurements (see
expanded scale in Figure 10b).

56
6. Poroelasticity and Dispersion
(Continued)
• Upon loading, there is both elastic and inelastic deformation
occurring whereas upon unloading, the initial slope corresponds
to mostly elastic deformation.

57
6. Poroelasticity and Dispersion (Continued)

Figure 10a. Hydrostatic loading and


unloading cycles of a saturated,
uncemented Gulf of Mexico sand
(Zoback., 2007).
58
6. Poroelasticity and Dispersion (Continued)

Figure 10b. An expanded view of


several of the cycles shown in
Figure 10a (Zoback., 2007).
59
7. Viscous Deformation in Uncemented
Sands
• Although cemented sedimentary rocks tend to behave elastically
over a range of applied stresses (depending on their strength),
uncemented sands and immature shales tend to behave
viscously.
• In other words, they deform in a time-dependent manner, or
creep, as illustrated schematically in Figure 5.
• Chang, Moos et al. (1997) discussed viscoelastic deformation of
the Wilmington sand utilizing a laboratory test illustrated in
Figure 11a.

60
7. Viscous Deformation in Uncemented
Sands (Continued)
• Dry samples were subjected to discrete steps of hydrostatic
confining pressure.
• After several pressure steps were applied, loading was stopped
and the sample was allowed to creep.

61
7. Viscous Deformation in Uncemented
Sands (Continued)
• After this creeping period, the sample was partially unloaded,
then reloaded to a higher pressure level and allowed to creep
again.
• Note that after each loading step, the sample continued to strain
at constant confining pressure and there was appreciable
permanent deformation at the conclusion of an experiment.

62
7. Viscous Deformation in Uncemented
Sands (Continued)
• One point of note that will be important when we consider
viscoplastic compaction of reservoirs that viscous effects are not
seen until the pressure exceeds the highest pressure previously
experienced by a sample.
• In other words, viscous compaction will only be important when
depletion results in an effective stress that is greater than the
sample experienced in situ.

63
7. Viscous Deformation in Uncemented
Sands (Continued)
• However, as will be shown below, once the previous highest
load experienced has been exceeded, viscoplastic compaction
can be quite appreciable.
• One would dramatically underpredict reservoir compaction from
laboratory experiments on uncemented sands if one were to
neglect viscoplastic effects.

64
7. Viscous Deformation in Uncemented Sands
(Continued)

Figure 11a. Drained hydrostatic load


cycling cleaned and dried Wilmington
sand (Zoback., 2007).
65
7. Viscous Deformation in Uncemented
Sands (Continued)
• In an attempt to understand the physical mechanism responsible
for the creep in these samples, Figure 11b illustrates an
experiment by Chang, Moos et al. (1997) that compares the
time-dependent strain of Wilmington sand (both dry and
saturated) with Ottawa sand, a commercially available
laboratory standard that consists of pure, well-rounded quartz
grains.

66
7. Viscous Deformation in Uncemented
Sands (Continued)
• Note that in both the dry and saturated Wilmington sand
samples, a 5 MPa pressure step at 30 MPa confining pressure
results in a creep strain of 2% after 2 × 104 sec (5.5 hours).
• In the pure Ottawa samples, very little creep strain is observed,
but when 5% and 10% montmorillonite was added, respectively,
appreciably more creep strain occurred.

67
7. Viscous Deformation in Uncemented
Sands (Continued)
• Thus, the fact that the grains in the sample were uncemented to
each other allowed the creep to occur.
• The presence of montmorillonite clay enhances this behavior.
• The Wilmington samples are composed of∼20% quartz,∼20%
feldspar, 20% crushed metamorphic rocks, 20% mica and 10%
clay.
• Presumably both the clay and mica contribute to the creep in the
Wilmington sand.

68
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 11b. Creep stress of Wilmington sand and Ottawa sand


(Zoback., 2007).
69
7. Viscous Deformation in Uncemented
Sands (Continued)
• Figure 12 illustrates a set of experiments that illustrate just how
important creep strain is in this type of reservoir sand.
• Note that after the initial loading step to 10 MPa, the creep
strain that follows each loading step is comparable in magnitude
to the strain that occurs instantaneously (Figure 12a).
• The cumulative strain (Figure 12b) demonstrates that the creep
strain accumulates linearly with pressure.

70
7. Viscous Deformation in Uncemented Sands
(Continued)

Figure 12a. Instaneous vs time


dependent strain dry Wilmington sand
(Zoback., 2007).

71
7. Viscous Deformation in Uncemented Sands
(Continued)

Figure 12b. Cummulative instantaeous


and creep strain dry Wilmington sand
(Zoback., 2007).
72
7. Viscous Deformation in Uncemented
Sands (Continued)
• Figure 13 summarizes four different ways in which viscoelastic
deformation manifests itself in laboratory testing, and
presumably in nature.
• As already noted, a viscous material strains as a function of time
in response to an applied stress (Figure 13a), and differential
stress relaxes at constant strain (Figure 13b).

73
7. Viscous Deformation in Uncemented
Sands (Continued)
• In addition, the elastic moduli are frequency dependent (the
seismic velocity of the formation is said to be dispersive) and
there is marked inelastic attenuation.
• Q is defined as the seismic quality factor such that inelastic
attenuation is defined as Q−1 (Figure 13c).
• Finally, a stress–strain test (such as illustrated in Figure 2) is
dependent on strain rate (Figure 13d) such that the material
seems to be both stiffer and stronger when deformed at higher
rates.

74
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 13a. Time-dependent deformation in a viscoelastic material is


most commonly observed as creep strain (Zoback., 2007).
75
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 13b. Time-dependent deformation in a viscoelastic material is


most commonly observed as stress relaxation (Zoback., 2007).
76
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 13c. Modulus dispersion and attenuation graph (Zoback.,


2007).
77
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 13d. Rate hardening graph (Zoback., 2007).


78
7. Viscous Deformation in Uncemented
Sands (Continued)
• The type of behavior schematically illustrated in Figure 13a is
shown for Wilmington sand in Figure 11a,b and the type of
behavior schematically illustrated in Figure 13b (stress
relaxation at constant strain) is shown for Wilmington sand in
Figure 14a.
• A sample was loaded hydrostatically to 3 MPa before an
additional axial stress of 27 MPa was applied to the sample in a
conventional triaxial apparatus.

79
7. Viscous Deformation in Uncemented
Sands (Continued)
• After loading, the length of the sample (the axial strain) was
kept constant.
• Note that as a result of creep, the axial stress relaxed from 30
MPa to 10 MPa over a period of ∼10 hours.
• An implication of this behavior for unconsolidated sand
reservoirs in situ is that very small differences between principal
stresses are likely to exist.
• Even in an area of active tectonic activity, applied horizontal
forces will dissipate due to creep in unconsolidated formations.
80
7. Viscous Deformation in Uncemented Sands
(Continued)

Figure 14a. Experiments on dry


Wilmington sand that illustrate
Relaxation of stress after
application of a constant strain
step (Zoback., 2007).

81
7. Viscous Deformation in Uncemented
Sands (Continued)
• The type of viscous behavior is illustrated schematically in
Figure 13d, and the rate dependence of the stress–strain
behavior is illustrated in Figure 14b for Wilmington sand.
• As expected, the sample is stiffer at a confining pressure of 50
MPa than it is at 15 MPa and at each confining pressure, the
samples are stiffer and stronger at a strain rate of 10 −5 sec−1, than
at 10−7 sec−1.

82
7. Viscous Deformation in Uncemented Sands (Continued)

Figure 14b. Experiments on dry Wilmington sand that illustrate The


strain rate dependence of sample stiffness (Zoback., 2007). 83
8. Thermoporoelastic
• Thermoporoelastic theory considers the effects of both pore
fluids and temperature changes on the mechanical behavior of
rock, it could be utilized as a generalized theory that might be
applied generally to geomechanical problems.
• Fundamentally, thermoporoelastic theory allows one to consider
the effect of temperature changes on stress and strain.

84
8. Thermoporoelastic (Continued)
• To consider the effect of temperature on stress Equation 12 is
the equivalent of Equations 11 where the final term represents
the manner in which a temperature change, ΔT, induces stress in
a poroelastic body:

..............(12)

85
8. Thermoporoelastic (Continued)
Sij = Stress tensor, psi
λ = Lame constant, psi
ε00 = Volumetric strain, volume
P0 = Pore pressure, psi.

86
8. Thermoporoelastic (Continued)
δij = Kronecker delta (δij = 1, i=j. δij = 0, i ≠ j).
εij = Strain tensor, volume
G = Shear modulus, psi
K = Bulk modulus, volume.

87
8. Thermoporoelastic (Continued)
αT = Coefficient of linear thermal expansion (α T = 1δL / LδT)
δL= Change in length of a sample, ft
L = Length of the sample, ft
δT= Change in temperature of the sample, oF.

88
9. Conclusions
1. There are 4 generic types of constitutive laws for
homogeneous and isotropic materials : Elastic constitutive
laws, poroelastic constitutive laws, elastic-plastic constitutive
laws, viscoelastic constitutive laws.
2. The theory of elasticity typically in terms of infinitesimally
small deformations, In this case, no significant damage or
alteration of the rock results from an applied stress and the
assumption that stress and strain are linearly proportional and
fully reversible is likely to be valid.

89
9. Conclusions (Continued)
3. A number of factors can make a rock mass anisotropic :
aligned microcracks, high differential stress, aligned minerals
(such as mica and clay) along bedding planes, macroscopic
fractures and faults.
4. Elastic anisotropy can have considerable effects on seismic
wave velocities, and is especially important with respect to
shear wave propagation.

90
9. Conclusions (Continued)
5. The three principal assumptions associated with theory are
similar to those used for defining pore pressure : First, there is
an interconnected pore system uniformly saturated with fluid,
Second, the total volume of the pore system is small compared
to the volume of the rock as a whole. Third, we consider
pressure in the pores, the total stress acting on the rock
externally and the stresses acting on individual grains in terms
of statistically averaged uniform values.

91
9. Conclusions (Continued)
6. The stiffness (elastic moduli) of a poroelastic rock is rate
dependent. In regard to seismic wave propagation, this means
that P-wave and S-wave velocities will be frequency
dependent.
7. At low confining pressure, there is about a factor of 2
difference between the moduli determined the two different
ways. As confining pressure increases the difference increases
significantly. Thus, there can be significant differences in
velocity (or the elastic modulus) depending on the frequency
of seismic waves.
92
9. Conclusions (Continued)
8. Although cemented sedimentary rocks tend to behave
elastically over a range of applied stresses (depending on their
strength), uncemented sands and immature shales tend to
behave viscously.
9. Because thermoporoelastic theory considers the effects of both
pore fluids and temperature changes on the mechanical
behavior of rock, it could be utilized as a generalized theory
that might be applied generally to geomechanical problems.
10.Fundamentally, thermoporoelastic theory allows one to
consider the effect of temperature changes on stress and strain.
93
Refrences
• Zoback, M.D., (2007) Reservoir Geomechanics. Edinburgh,
Cambridge : Cambridge University Press, Chapter 3.

94

You might also like