Sei sulla pagina 1di 13

ARMA 16-496

Where does the stress path lead?


Irreversibility and hysteresis in reservoir geomechanics
Holt, R.M.
NTNU, Trondheim, Norway
Gheibi, S.
NTNU, Trondheim, Norway
Lavrov, A.
SINTEF Petroleum Research, Trondheim, Norway

Copyright 2016 ARMA, American Rock Mechanics Association


This paper was prepared for presentation at the 50th US Rock Mechanics / Geomechanics Symposium held in Houston, Texas, USA, 26-29 June
2016. This paper was selected for presentation at the symposium by an ARMA Technical Program Committee based on a technical and critical
review of the paper by a minimum of two technical reviewers. The material, as presented, does not necessarily reflect any position of ARMA, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written consent
of ARMA is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 200 words; illustrations may not be copied. The
abstract must contain conspicuous acknowledgement of where and by whom the paper was presented.
ABSTRACT: Reservoir stress changes caused by pore pressure depletion have impacts on hydrocarbon recovery, on safety and
stability margins during drilling, and on possible seismicity. Stress path prediction is a challenge, in particular when pore pressure
is increased by injection after depletion, either for improved recovery or for storage of e.g. CO2. Three different perspectives of
stress path reversal are considered here: i) The mechanical behavior of rock that undergoes unloading and reloading, starting from a
virgin state, has been addressed through controlled laboratory experiments, ii) An analytical approach has been developed to sketch
the interplay of tensile and shear fractures with the stress path itself, and iii) Numerical modeling has been performed to account for
stress heterogeneities and plastic deformation of the reservoir rock. The results show that all these mechanisms may contribute to
the resulting stress path, in addition to the role of thermal stresses associated with injection of cold fluids into reservoirs.

Stress arching, associated with v > 0 , is enhanced by


1. INTRODUCTION
increasing aspect ratio (= thickness divided by diameter
The concept of reservoir stress path is becoming a more of the depleting zone), by increasing stiffness of the
and more relevant tool for describing and understanding surrounding rock relative to the reservoir rock, and by
the dynamic behavior of hydrocarbon or storage reservoir tilt (see Fjr et al., 2008, Ch. 12).
reservoirs. Reservoir stress path coefficients are defined
(Hettema et al., 2000) as the change in the total stress The importance of stress path was traditionally seen in
per unit change in the pore pressure Pp (subscripts "v" prediction of compaction and subsidence and potential
and "h" denote vertical and horizontal directions): fault activation during depletion (Zoback and Zinke,
2002; Chan et al., 2004). It has become clear that the
v h stress path inside the reservoir as well as in the
=v = ; h (1)
Pp Pp overburden has a profound influence on the ability to
drill safe and stable wells into a reservoir that is being or
v and h are often called arching coefficient and has been depleted. In the permeable zones of the
depletion coefficient, respectively. In case of a reservoir, the horizontal stress drops as pore pressure
laterally infinite reservoir in a linearly elastic and decreases, as illustrated in Eq. (1), leading to increased
homogeneous subsurface, no stress is transferred to the risk of mud losses: While the value of the lost-
overburden, v = 0 , and the reservoir is deforming circulation pressure in a naturally-fractured reservoir
depends largely on the orientation and morphology of
uniaxially (no horizontal strain). The depletion natural fractures (Lavrov, 2016a), its value in the
coefficient is then directed by the Poisson's ratio of the absence of natural fractures depends on the far-field
drained reservoir rock, and by the Biot coefficient : minimum horizontal stress, h . The value of h (with
1 2 some safety margin) is commonly used as the upper
h = (2)
1 bound of the pressure window during overbalanced
drilling and during well cementing (Lavrov, 2016b).
Since the total minimum horizontal stress decreases presented, pointing to the importance of spatial variation
during depletion, the risk of lost circulation in a depleted of stress path during depletion and injection, associated
reservoir is greater than it was before production started. with accumulation of localized plastic deformation.
Reducing mud weight to cope with such scenarios may
however trigger borehole collapse in the non-permeable
parts of the open hole section, leading to stuck pipe and 2. LABORATORY SIMULATIONS USING
increased drilling costs. If the stress path coefficient SYNTHETIC SANDSTONE
during subsequent water injection is smaller than during
Several experimental studies have been performed in our
depletion, the lost-circulation pressure will not be
laboratory, utilizing synthetic sandstone formed under
restored to its pre-depletion level at the end of injection.
stress. The main purpose of these studies was to compare
The stress path during injection following depletion was compaction under in-situ conditions with compaction of
discussed by Santarelli et al. (1998) based on field data a stress-relieved and reloaded core sample (Holt and
from a poorly consolidated sandstone reservoir in the Kenter, 1992; Holt et al., 2000; Holt and Stenebrten,
North Sea, supported by a few other observations. They 2013; Holt et al., 2014). In these tests, repeated load
found that the stress path during repressurization by cycles were applied to the material, enabling a
water injection was essentially the same as that systematic study of how the stress path would change
measured after maximum depletion, corresponding to a because of altered material properties, albeit still
value of h being close to zero during injection. This following the same uniaxial strain path.
would correspond to a Poisson's ratio of 0.5 according to
These laboratory tests have been performed on dry
an elastic model (Eq. (2)). Santarelli et al. speculate that
material, so that the total axial and radial stresses ( z
this behavior is a result of large-scale rock mechanics,
somehow leading to a rigid but perfectly plastic and r ) would mimic the effective vertical and
behavior. They point out that core measurements in the horizontal stresses in the field ( = v v Pp and
laboratory do not show similar results, indirectly
h h Pp , respectively). In the case of an infinitely
=
attributing the field scale behavior to faults and fractures
beyond the laboratory scale. They also refer to wide reservoir (no stress arching), the effective stress
temperature changes in the injection zone as a possible changes in the reservoir are (using the effective stress
explanation. In a more recent paper (Santarelli et al., principle and Eq. (1)) v =Pp and h = (1 h )v .
2008), the role of temperature was further highlighted in
a study addressing effects of long-term water injection in Thus, the depletion coefficient when computed from
a North Sea field complex. laboratory data is given by the Biot coefficient and the
ratio r z =
K0 :
In this paper, we present three different studies of
possible mechanisms of stress reversal when a depletion r
phase is followed by injection. This could be the result h =(1 )=(1 K 0 ) (3)
of an enhanced oil recovery operation by e.g. water z
injection, or it could be the situation when injecting CO2 Several laboratory experiments have been analyzed.
into a depleted reservoir. The altered stress path would Figure 1 shows the calculated depletion coefficient based
clearly affect infill drilling operations, increasing the risk on uniaxial compaction (loading, representing simulated
of mud losses. A key question in several places is to depletion) with subsequent uniaxial extension
what extent the injection process may trigger (unloading, representing simulated injection) with a very
earthquakes: Although releasing stresses normally leads soft synthetic sandstone (hereafter termed Soft). This
to reduced acoustic emission (AE) activity (Kaiser material, having approximately 30 % porosity, was
effect) (Lavrov, 2003; Becker et al., 2010) in the formed by using sodium silicate solution as a bonding
laboratory and reduced seismicity in field situations, agent. The sample was cemented with amorphous silica
there may be a finite risk that critically oriented faults bonds created when flushing the sample with CO2 at an
may be activated. axial stress of 7 MPa and a confining pressure of 3.5
We start by showing results from controlled laboratory MPa. The sample was left to stabilize prior to loading as
tests, addressing stress path hysteresis in studies of well as unloading in order to minimize creep effects.
synthetic rocks formed under stress. Clearly, such Table 1 shows Youngs modulus and Poissons ratio for
studies are idealized and provide only partly insight into the Soft sandstone, and the Biot coefficient based on an
the field problem. Faults have to be accounted for, and anticipated solid grain modulus of 36 GPa (almost pure
here we present an analytical approach to describe the quartz). From Figure 1, the depletion coefficient is seen
hysteresis related to tensile and shear fractures created or to decrease with increasing axial stress (mimicking
activated during depletion and subsequent injection. depletion), but it jumps to a higher level upon stress
Finally, numerical simulations on the reservoir scale are reversal (mimicking injection). This is associated with a
reduction in Poissons ratio (increasing K 0 ), which agent. No stabilization period was used, which led the
sample to continue to contract slightly during initial
more than compensates for the reduction in caused by
unloading (simulating injection). Hence, the stiffness
stiffening of the material during unloading (cf. Eq. (3)).
during unloading is likely to be overestimated. The
The stress path coefficient then stays constant until the
figure shows that reloading after the unloading
sample is unloaded past the previous stress, after which
(injection) segment more or less leads to the same stress
h starts to decrease. This is evidence of stress memory path as during initial compaction (depletion). However,
(Kaiser Effect), and was also seen in acoustic emission
the compaction modulus has increased (reducing ) and
recordings during an identical test with a similarly
Poissons ratio has decreased (decreasing K 0 ), so that
manufactured sample. The Figure contains a calculated
dynamic depletion coefficient (as seen in Table 2), the product (1 K 0 ) remains almost constant.
derived from P- and S-wave velocities measured during
the simulated depletion experiment. The dynamic
coefficient is close to the static at the very beginning of
the experiment, in line with a previously noticed
observation in such experiments (Holt et al., 2013). It
does not however show the decreasing trend of the static
depletion coefficient, while the dynamic and static
values coincide during initial unloading, in agreement
with observations and theoretical arguments relating the
difference between static and dynamic behavior to non-
elastic static deformation (Fjr et al., 2013). Similar
results are also seen in other soft synthetic sandstones.

Figure 2. Depletion coefficient calculated on basis of uniaxial


compaction and extension of Stiff synthetic sandstone,
formed at 30 MPa axial stress and 15 MPa confining pressure.
The figure includes static h as well as dynamic h calculated
from ultrasonic measurements during the test. Notice the
possible influence of creep during unloading (simulated
injection), which is why the reloading behavior is also shown
in this Figure.

Experiments with these and similar materials show that


the hysteresis in Poissons ratio during the 1st loading-
unloading cycle becomes smaller the more competent
Figure 1. Depletion coefficient calculated on basis of uniaxial
the rock is, while is always reduced by the increasing
compaction and extension of Soft synthetic sandstone,
formed at 7 MPa axial stress and 3.5 MPa confining pressure. bulk modulus. Thus, the stress path may decrease upon
The figure includes static h as well as dynamic h calculated going from loading to unloading in competent rocks,
from ultrasonic measurements during the test. even in the absence of creep. The observations largely
confirm, as was seen for the Soft synthetic sandstone,
that incorporating the dynamic moduli in
Experiments were also performed with a competent
porous material (hereafter termed Stiff) made by
dyn (1 K 0,dyn ) (Eq. (1)) yields a good estimate of the
gluing sand grains with epoxy at an axial stress of 30 reversed stress path.
MPa and 15 MPa confining pressure, with a porosity of As mentioned, the laboratory experiments are only valid
around 20 %. Data for this material are shown in Table 1 simulations of the field situation for the case of uniaxial
and Table 2 (because of the epoxy content, the solid strain, i.e. no stress arching. One may extrapolate the
grain modulus was anticipated to be lower than for laboratory findings towards a field situation by using a
quartz and somewhat arbitrarily set to 32 GPa). geomechanical model that permits finite reservoir size in
Figure 2 shows that the depletion coefficient remains addition to elastic contrast between the reservoir and the
almost constant and very similar to its dynamic overburden. Based on analytical relationships derived by
counterpart during initial loading (simulating depletion). Mahi (2003) from Finite Element (FEM) simulations by
This particular synthetic rock experienced creep, Mulders (2003), these effects were incorporated for the
probably related to the ductile nature of the bonding soft and the stiff synthetic sandstones described above.
In both cases the ratio between the height and the Table 3 shows that for the Soft sandstone, the depletion
diameter (the aspect ratio) of the depleted/inflated zone constant h always increases during the rebound, also
was set to 0.1. Both synthetic sandstones were placed in when arching is present. For the Stiff sandstone, there is
elastically matched surroundings. In addition, the Soft always a slight decrease. Stiff surroundings promote
sandstone was placed in stiff surrounding rock with E = arching, as expected, whereas arching is reduced in the
5 GPa and = 0.2, and within soft surrounding rock with case of soft surroundings.
E = 1 GPa and = 0.4. The Stiff sandstone was placed in
stiff surrounding rock with E = 30 GPa and = 0.18, and Shahri and Miska (2014) performed pore pressure
within soft surrounding rock with E = 1 GPa and = 0.4. depletion (loading) experiments with liquid-saturated
Berea sandstone, also investigating the effects of
Table 1. The tangent moduli (Youngs modulus E, Poissons subsequent pressure build-up (unloading). They found
ratio , and Biot coefficients ) measured during loading and that the stress path decreased upon stress reversal, and
unloading of the Soft and Stiff synthetic sandstones. The argued that this was caused by an increase in Poissons
measurements are made at different stress levels z above the ratio. In our samples, we do not observe such an
forming stress (see captions of Figures 1 and 2). increase, and the decrease in stress path for competent
Eload / Eunl load / unl load/ unl
rock is attributed to reduction of the Biot coefficient.
z There are however two important differences that make
[MPa] these tests difficult to compare: Saturated and dry rocks
[GPa] [-] [-] may behave differently, and controlling stress history
SOFT 4 0.8 / 2.2 0.34 / 0.29 0.98/0.95 such as in our experiments with synthetic rocks may
make a difference as compared to working with cored
STIFF 30 13.5 / 21.2 0.29 / 0.29 0.67/0.48 materials that are stress-released.
It is clear that the significant decrease in depletion
Table 2. Static and dynamic depletion coefficients estimated constant seen by Santarelli et al. (1998) cannot be
from laboratory data for Soft and Stiff synthetic sandstones, at explained by purely elastic arguments. As hinted in their
different stress levels z above the forming stress (see
1998 paper and elaborated further by Santarelli et al.
captions of Figures 1 and 2, and Table 1).
(2008), thermal stress effects may largely explain
z h,load h_unl h_dyn irreversible stress paths like those observed in the field,
[MPa] [-] [-] [-] if the reservoir has been cooled by continuous water
injection. Hettema et al. (2004) reported several cases of
SOFT 4 0.47 0.57 0.60
lost circulation in North Sea wells drilled close to
STIFF 30 0.40 0.29 0.40 injector wells, and found the horizontal stress to be 5-10
MPa lower than expected from depletion alone. Sreide
et al. (2014) carried out laboratory tests on reservoir
Table 3. Depletion coefficient h and arching coefficient h sandstone from offshore Norway. Cooling from 140 to
estimated on basis of numerical FEM modeling (Mahi, 2003), 60 C led to 14-22 MPa decrease in confining pressure
using the laboratory data obtained with Soft and Stiff synthetic for samples held in uniaxial strain conditions with an
sandstones. The Table lists the influence of the surrounding applied initial in-situ vertical (= axial) stress of 45 MPa.
rock properties; being stiff (E = 30 GPa and = 0.18) or soft The thermal expansion constant th was in the range 1-
(E = 1 GPa and = 0.4).
1.310-5 C-1. The thermal effect can be incorporated
SOFT h,depl v,depl h,inj v,inj according to Perkins and Gonzales (1984), who showed
Sst Reservoir that for a cooled reservoir with small height-to-diameter
ratio, the horizontal stress reduction is given by
Matched 0.47 0.04 0.57 0.05 E
thh th
= T (4)
Stiff surroundings 0.52 0.19 0.59 0.10 1
Soft surroundings 0.48 0.05 0.54 0.03 Here T is the difference between reservoir and
h,depl v,depl h,inj v,inj surrounding rock temperature, i.e. negative for cooling.
STIFF
For the synthetic rocks above, the thermal effect
Sst Reservoir
amounts to 0.01-0.05 MPa/C for the Soft and 0.2-0.4
MPa/C for the Stiff sandstone, provided there is minor
Matched 0.58 0.06 0.56 0.05
difference between their thermal expansion coefficients.
Stiff surroundings 0.61 0.10 0.58 0.07 Hence, one would expect a significant thermal effect for
a stiff rock. With large temperature changes, the effect
Soft surroundings 0.27 0.02 0.26 0.02
should also be noticeable for soft rocks.
vertical dimension. Thus, the stress arching effect is
neglected here.
3. STRESS PATH AND IRREVERSIBLE
FRACTURE PERMEABILITY: TENSION During depletion, the vertical total stress in such a
FRACTURES reservoir is unchanged and remains equal to the stress
created by overburden's bulk weight. The horizontal total
Most rocks do contain fractures. Let us consider the stress normal to the fracture, h , decreases during
effect that stress changes during production and
depletion might have on tension (mode-I) fractures depletion. The effective stress normal to the fracture,
oriented normally to the minimum horizontal in-situ h , increases during depletion. As a result, the fracture
stress, h (normal faulting (extensional) tectonic regime closes, and its permeability decreases (line 1-2 in Figure
is assumed in this Section and the next Section). 3).
Fractures normal to h are important for production During injection, the total vertical stress remains
from at least some naturally-fractured reservoirs unchanged, while the effective vertical stress decreases.
(Lorenz, 1999) since, all the rest being equal, fractures As a result, the effective horizontal stress decreases
of this orientation have the greatest apertures. (Lavrov, 2016c). The extent of this decrease depends on
the horizontal stress path coefficient, h . If the stress
The permeability of a fracture under normal stress
exhibits nonlinearity and hysteresis, as schematically path coefficient is the same during injection as it was
shown in Figure 3. Similar behavior is characteristic of during depletion, restoring the reservoir pressure to the
rock matrix permeability under loading/unloading pre-depletion level will restore the effective horizontal
cycles, where it is due to irreversible closure of pores stress in the reservoir to its pre-depletion value (point 3
and microcracks (Lavrov, 2005). In the case of a single in Figure 3). If the stress path coefficient during
tension fracture, this behavior is due to plastic injection is smaller than during depletion, the effective
deformation and crushing of asperities on the fracture horizontal stress in the reservoir will be smaller after
faces under compression, which leads to irrecoverable injection than it was before depletion (point 4 in Figure
permeability damage. 3). In any event, the hysteretic behavior of the fracture
permeability makes it highly unlikely that the fracture
permeability after injection will be equal to that before
depletion. Thus, it is, in practice, hardly possible to
restore the permeability of a tension fracture to the pre-
k depletion level by re-pressurizing the reservoir to the
1
pre-depletion pore pressure value.
We saw above that the fracture permeability during
4
injection is influenced by the reservoir stress path
coefficient (point 3 vs. point 4 in Figure 3). The stress

3
path coefficient itself, however, is likely to be influenced
by the fractures. Consider again fractures normal to h .
2

The fracture aperture during the depletion-injection
cycle behaves similarly to the fracture permeability
(Figure 4).
As evident from Figure 4, the fracture is stiffer during
unloading (injection) than during loading (depletion). If
h we now consider a rock with many such fractures
Figure 3. Nonlinearity and hysteresis of fracture permeability (Figure 5), the behavior of the rock mass will be affected
under normal loading. Fracture permeability, k, as a function by the individual fractures' behavior. As a result, the
of normal effective stress, h . 1 - pre-depletion state; 2 effective stiffness of the rock mass in the direction of
depleted reservoir; 3 after injection, if stress path during h will be higher during injection than during depletion.
injection is the same as during depletion; 4 after injection, if The deformability in that direction will be, accordingly,
stress path during injection is smaller than during depletion.
lower during injection than during depletion. This is
equivalent to having the Poisson's ratio for horizontal
Let us consider how the permeability of such fractures deformation lower during injection than during
changes during depletion of and subsequent injection depletion. Consequently, the irreversible normal
into the reservoir. It is assumed that the horizontal deformation of tension fractures acts to increase the
dimensions of the reservoir are much larger than its
stress path coefficient h during injection, as compared
to the h values during depletion. The extent to which Let us analyze how slip of a shear fracture may develop
during depletion and subsequent injection. The fracture
this effect can influence h depends on the availability is assumed to be located in the reservoir and to be
and properties of tension fractures normal to h . As parallel to the intermediate in-situ stress, H (normal
pointed out by Smart et al. (2001), the dominant type of tectonic regime is assumed here, as before).
fractures in a reservoir are not tension fractures oriented
Before depletion, the fracture is assumed to be in
normal to h , but shear fractures that may have
equilibrium, i.e. the shear stress on its plane is not
different orientations. (It is also quite unlikely that a sufficient to overcome Coulomb's friction. During
fracture set in situ is oriented exactly normal to the depletion, the total vertical stress remains constant,
currently acting h , since fractures develop during the while the total horizontal stress decreases. In the normal
entire geological history, and their orientations thus need tectonic environment, this would imply that the shear
not be related to the present-day stress directions.) stress on the fracture increases. Counteracting this effect
is the reduction of the pore pressure, which leads to an
w increase of the effective normal stress acting on the
fracture surface. This leads to an increase in the
frictional resistance on the fracture plane. It is intuitively
clear that in order to induce shear (and, thus, alter the
Loading permeability) of a shear fracture under depletion, the
(depletion) stress path coefficient, h must be sufficiently high so
that the effect of reduction in h dominates. Let us
evaluate how high the stress path coefficient needs to be
in order for the net shear stress on the fracture plane
Unloading during depletion to be higher than before depletion. The
(injection) net shear stress is defined as follows:
net = n (5)
h
where is the shear stress on the fracture plane; is
Figure 4. Fracture aperture, w, as a function of normal the friction coefficient between the fracture faces; n is
effective stress, h . the effective normal stress on the fracture plane. The
cohesion between the fracture faces is neglected here,
v and the Biot effective stress coefficient used to compute
n will be assumed equal to unity in this Section (both
assumptions are common in fracture mechanics and will
h not affect the general conclusions drawn from our
analysis.)
Eq. (5) shows that the net stress being zero would not
imply that the shear stress has dropped to zero. It would
rather imply that the shear stress is balanced by the
frictional stress.
The net shear stress acting on the fracture plane before
depletion is given by:

Figure 5. Rock volume with a fracture set normal to h . =


01 (0)
(net)
2
(
v (h ) sin 2
0
)
(6)
(0) (0) (0) (0)
4. STRESS PATH AND IRREVERSIBLE 2
( 2
) (
v + h h - v cos 2 + Pp( )
0
)
FRACTURE PERMEABILITY: SHEAR (0) (0) (0)
FRACTURES where v , h , Pp are the vertical in-situ stress, the
As mentioned in the previous Section, shear (or minimum horizontal in-situ stress, and the pore pressure
combined shear and tension) fractures are more common in the reservoir before depletion, respectively; is the
in reservoirs than pure tension (mode-I) fractures. angle between the fracture plane and the vertical stress.
Consider now depletion of the reservoir from pore depletion ( (h ) = (h ) ) and (ii) the stress path coefficient
i d
(0) (d )
pressure Pp to pore pressure Pp < Pp(0) . If the stress (i )
during injection is equal to zero ( h = 0 ).
path coefficient during depletion is equal to (h ) , the net
d

If (h ) = (h ) and the reservoir is re-pressurized to the


i d
shear stress after depletion is given by:
original, pre-depletion pore pressure level, the pre-
(netd ) =
(net
0) 1 d
( )
+ (h ) Pp( ) Pp( ) sin 2
2
0 d depletion status quo (in terms of pore pressure and in-
situ stresses) will be restored at the end of the injection.
+ ( ) ( P ( ) P ( ) ) cos
h
d
p
0
p
d 2
(7) This will stabilize the shear fracture. There will be no
additional slip on the fracture plane during injection,
( P ( ) P ( ) ) and, subsequently, no further permeability change.
0 d
p p
(i )
If h = 0 , the total in-situ stresses do not change, while
(d ) (0)
Thus, for net to be greater than net , the stress path the pore pressure increases during injection. The shear
coefficient must meet the following condition during stress on the fracture plane thus does not change, while
depletion: the effective normal stress decreases, thereby increasing
the net shear stress. Thus, if slip was initiated on an

(h ) >
d
(8) oblique fracture during depletion, it will in this case
cos ( sin + cos ) continue during injection. If slip was not initiated during
depletion, it might or might not be initiated during
If this condition is not met, there will be no slip on the injection.
fracture during depletion if there was no slip before
depletion. In other words, if the condition given by Eq. This analysis shows that the stress path coefficient
(8) is not met, fracture will become more stable during controls whether or not slip will continue, will cease or
depletion. If the condition given by Eq. (8) is met, there will start on an oblique fracture during injection. In any
might be slip. Whether slip indeed occurs will depend on event, the slip created during depletion (and thus the
how close the fracture was to slip in the original (pre- fracture permeability alteration) cannot be reversed
depletion) state. If the in-situ stress anisotropy was large during subsequent injection.
enough, and the friction coefficient sufficiently small, Note that no assumptions about elasticity or plasticity of
slip may develop during depletion, provided Eq. (8) the rock have been made in this Section. To arrive at the
holds for the stress path. conclusion marked by italic, it is not necessary to invoke
Eq. (8) demonstrates that the stress path coefficient any specific mechanism of stress path change during the
controls whether or not slip may occur on an oblique depletion/injection cycle.
fracture during depletion. In addition to the effect of stress path on slip activation
Permeability of a fracture can be changed in different at shear fractures and faults discussed above, the slip
ways during slip. If the effective normal stress, n , is itself may affect the stress path. If a shear fracture (or a
fault) has been activated during depletion, the in-situ
sufficiently low, the fracture may experience dilation stresses in its vicinity will be no longer controlled by the
caused by asperities of the fracture faces. As a result, the properties of the bulk rock, but will instead be
fracture permeability may increase (Esaki et al., 1999). determined by the slip condition. This condition reads as
If slip continues, the permeability may again decrease follows:
because of gouge accumulation (Smart et al., 2001). If
the effective normal stress is sufficiently high, the (netd ) = 0 (9)
fracture permeability may start decreasing from the very
beginning of the slip (Gutierrez et al., 2000). The net shear stress on the fracture (or fault) during
depletion is given by:
Let us now consider what happens to the fracture during
subsequent injection into the reservoir. During injection,
the total vertical stress is unchanged, while the total
1 (d )
(netd )
=
2
( )
v (hd ) sin 2
horizontal stress increases (if the stress path coefficient
d d
during injection is greater than zero) or remains
unchanged (if the stress path coefficient is zero). 2
( ) (
2
d
)
(v ) + (h ) (h ) - (v ) cos 2 + Pp( ) (10)
d d

Consider two extreme cases: (i) the stress path =0


coefficient during injection is equal to that during
(d ) (d ) (d )
where v , h , Pp are the vertical in-situ stress, the heterogeneity, presence of joints and faults, boundary
conditions, etc. To investigate the stress path hysteresis
minimum horizontal in-situ stress, and the pore pressure on the field scale, a numerical model was set up in a
in the reservoir during depletion, respectively. hybrid FEM/DEM in-house code called MDEM (Alassi,
Assuming, as before, that the total vertical stress near an 2008; Lavrov et al., 2015). In the elastic regime, the
activated shear fracture remains unchanged during calculation is equivalent to FEM, and the problem
depletion (no stress arching), and taking the derivative of becomes that of a discontinuum when cracks develop in
Eq. (9) with respect to Pp( ) , we obtain the following
d the material. MDEM can also model plasticity
(softening-hardening) after the material reaches its yield
stress path coefficient for a shear fracture activated point. Therefore, it can handle the linear elastic,
during depletion: softening, and discrete behavior of materials. Moreover,
it can be coupled to fluid flow simulators. In the
d (hd )
= (11) examples in this paper, MDEM is not coupled to any
dPp (d )
cos ( sin + cos ) flow simulator, and the depletion or injection is
mimicked by changing the pore pressure in a defined
Thus, the activated shear fracture may affect the stress region of the domain. The simulations are performed
path in its vicinity. If many shear fractures (or faults) are with a 2D version of the model. The reservoir is a 1 km
activated in the reservoir during depletion, the reservoir x 100 m region with half-circle edges to smoothen the
stress path may be significantly affected. reservoir boundary and reduce the stress concentrations
There is thus a two-way coupling between shear at the corners (Figure 6).
fractures (faults) and the stress path during depletion: Table 4 shows the rock mechanical properties of the
the activated shear fractures affect the stress path, and reservoir (as I in the table) and the surroundings (as II in
the stress path governs the slip on these fractures. the table). The effective in-situ stresses are 50 and 20
During subsequent injection, the shear fractures may MPa, and the stress regime is normal faulting, meaning
become stabilized (irreversibility of slip discussed that the vertical stress is the maximum principal stress.
above). The reservoir stress path during injection will in The material can undergo plastic deformation, and the
this case be largely determined by the bulk properties of plasticity modulus is equal to 108 Pa. The plasticity
the reservoir rocks. Thus, the stress path coefficient may modulus represents the stress drop due to the softening
be different during injection and during depletion, if it is of the rock after peak strength. This low value was
influenced by active shear fractures and faults. The chosen to make the material deform in a ductile manner
magnitude of this difference will depend on the rather than to fail in a brittle fashion. The initial pore
properties of the fractures (inclination, abundance, pressure is assumed to be zero.
friction coefficient) and the anisotropy of the far-field in-
situ stresses.
It should be noted that, in reality, a fault or a shear
fracture might be activated only at some locations
(Longuemare et al., 2002). This will further complicate
its effect on the reservoir stress path.
Our analysis leading to Eq. (11) is inspired by the
similar analysis performed by Addis et al. (1996) for a
fault of the most critical orientation. Our analysis is an
extension of their work to faults and fractures of an
arbitrary orientation since, as pointed out also by Addis
et al., the orientation of faults and fractures does not
need to be linked to the contemporary in-situ stress field.
Our Eq. (11) reduces to Addis et al.'s Eq. (9) for the
Figure 6. The 1km x 100 m reservoir model under a normal
most-critically stress fault.
faulting stress regime.

To simulate depletion, the pore pressure was reduced


5. NUMERICAL MODELING OF FIELD-
from 0 to -32.5 MPa. As a result of the pore pressure
SCALE STRESS PATH HYSTERESIS reduction, the material undergoes plastic deformation.
Rock behavior is scale dependent, and laboratory scale Therefore, the stress redistribution in the subsequent
behavior might not necessarily represent the rock simulation steps is path dependent.
behavior on the field scale. This may be due to
Table 4. Mechanical properties of the rock in the numerical and late depletion stages, respectively. It is clear from
model. Figure 8 that the stress path coefficient is different for
Tensile Shear Softening the O => A and A => B cases, and it decreases as the
E Friction depletion continues. This is because of the damage
strength strength modulus
(GPa) coefficient (plastic deformation) that occurs in the model. The
(MPa) (MPa) (Pa)
I 15 0.3 1.6 4-4.7 0.35 1e8 damage accumulates in the center of the reservoir first,
II 15 0.3 1.6 4-4.7 0.57 1e8 then due to greater reduction in the pore pressure and the
increase in the vertical stress at the outer flanks, the
number of the damaged elements increases in the outer
To reduce the path dependency, pore pressure was
flanks. In the center of the reservoir (-100 m < x < 100
changed by 1 MPa per step. After the depletion had
m) the stress path fluctuates at A => B, but generally it is
reached the final pressure (-32.5 MPa), the pore pressure
lower than the stress path at O => A, because of the
was increased in 1 MPa steps to simulate the injection.
damage incurred in the reservoir center. However, the
Figure 7 indicates the different depletion and injection
reduction in the stress path is significant in the other
stages with their names and specifies the amount of
parts of the model, i.e. at the inner and outer flanks. It
pressure change at each stage. There are two scenarios:
even becomes negative for the outer flanks, meaning that
only injection and injection after depletion. The
the depletion will increase the x-stress at the outer
only injection scenario is from point O (the initial
flanks, which is opposite to the elastic solution. The
state) to G in Figure 7, with an accompanying pore
reason is the damage accumulation near the outer flanks.
pressure increase of 5.5 MPa. The injection after
depletion scenario proceeds from O to A and point B by After the depletion, the injection was simulated by
decreasing the pore pressure to -32.5 MPa in 1 MPa increasing the pore pressure. As seen in Figure 8a, the
steps. In O => A stage, the pore pressure was decreased stress path (after final depletion) is constant up to +16
by 0.5 MPa to obtain the stress path in the very MPa pore pressure increase (step C, net pressure is -16.5
beginning of the depletion where the model is in the MPa = -32.5 MPa + 16 MPa). However, after this stage,
elastic domain (Figure 8a). additional damage was created at the inner flanks and the
stress path changed. After the failure in the inner flanks
The injection stages are began from point C to F with 16,
during pressure rebound, the stress path coefficient
32, 37.5 and 45 MPa pressure increase with respect to
increased. The maximum increase was recorded in the
point B, i.e. the final depletion stage. The colors of the
inner flanks. The amount of damage (number of
markers in Figure 7 correspond to the Mohr circles in
damaged elements) inside the reservoir increased by a
Figure 8 and Figure 9.
factor of three compared to the previous (when it was
mainly in the inner flanks). Figure 8b, c and d indicate
the most unstable Mohr circles for different depletion-
injection stages at the center, inner and outer flanks
respectively. For the reservoir center, the green Mohr
circle indicates the stress state after -32.5 MPa depletion
and 32 MPa pressure rebound, meaning that the
reservoir pressure has almost fully rebounded. If there
were no hysteresis, the circle would return to the initial
stress state (the dark blue circle). However, we recorded
a hysteresis which is not very significant for the
reservoir center. If the injection is continued up to the
net pressure of +13.5 MPa (magenta colored circle), the
Mohr circle at the reservoir center moves towards the
failure envelope in Figure 8. The dashed lines in
Figure 7. The names of the different cases with pressure Figures 8 and 9 represent failure envelopes for possible
change information in each case. pre-existing faults or fractures in the reservoir or the
surrounding rocks. In contrast to the reservoir center, the
Results are shown in the form of depletion coefficient hysteresis is significant in the inner and outer flanks
(Figure 8c, d), especially in the inner flank. The rebound
( h ) along a horizontal line passing through the
stress state is more unstable compared to the initial stress
reservoir center (Y=0) and plots of the most unstable state. Increasing the pore pressure further might be
Mohr circles corresponding to the reservoir center, inner dangerous.
and outer flanks for different cases. Figure 8-a shows the
horizontal stress path along the line Y=0. The curves It is also important to investigate the differences between
O => A and A => B show the stress paths for the early the injection into a depleted field and into an intact
formation, such as during CO2 injection into a saline The numerical simulations presented above demonstrate
aquifer. Figure 9 indicates the stress paths and the Mohr that plastic deformation has a significant impact on the
circles corresponding to the reservoir center, inner and stress path: h decreases during depletion, with the
outer flanks for the two scenarios of only injecting up to amount of the stress path reduction depending on the
5.5 MPa (case G in Figure 7) and pressurizing by 37.5 location within the reservoir. The hysteresis might be
MPa after depletion by 32.5 MPa (E in Figure 7; , final
pressure +5.5 MPa). As discussed above, the hysteresis negligible in the central parts of the reservoir, but h
in negligible in the center, however, there is a decreases considerably in the inner and outer flanks in
considerable hysteresis in outer and inner flanks. The our simulations. Injection was seen to change the stress
inner flank (Figure 9c) for both scenarios might be state differently, causing plastic deformation in different
unstable. However, the injection into the depleted model locations of the reservoir and its surroundings. In our
is significantly more unstable. Also, in the pure injection simulations, the inner flanks failed, and h increased in
case, the outer flank becomes stable, while it may the flanks as a result. Clearly, the resulting stress path is
become unstable during injection into the depleted different if injection is started without preceding
reservoir. depletion.
So far, we have observed that the stress path hysteresis is The presence of faults or fractures is known to affect the
greater in the outer and inner flanks than in the center of stress path (Addis et al., 1996). Here, an analytical
the reservoir. Figure 10 indicates the variation of the approach has been presented, showing that the stress
stress path with increasing number of damaged elements path coefficient in a fractured reservoir is affected by
at locations of -450 m and -470 m in the reservoir flanks. irreversible behavior of tensile and shear fractures
According to Figure 10, the stress path decreases slightly (faults). On the other hand, the stress path itself affects
by increasing damage at the center due to depletion. The reactivation of shear fractures and faults during depletion
stress path reduction becomes significant after -19.5 and injection. As a result, it affects microseismicity and
MPa pore pressure reduction. This is due to the failure at permeability evolution in a naturally-fractured and
the outer flanks. In the rebound, there is no significant faulted rock. There is thus a two-way coupling between
change in the stress path up to the 23 MPa pressure the fracture deformation and the stress path coefficient.
rebound. However, if the injection is continued, the
inner flank is damaged, and the stress path increases at Permeability caused by tensile fractures exhibits
both locations and becomes constant at higher injection irreversible, hysteretic behavior during
pressures. depletion/injection for a naturally-fractured reservoir. As
a result, it is not possible to restore the permeability of
such a reservoir to the pre-depletion level by simply
6. CONCLUSIONS repressurizing the reservoir, even if the same stress paths
are followed during production and injection.
The title of this paper indicates that the authors aim at
giving clear directions as to where the stress path leads. The final permeability of tensile fractures after injection
The work performed shows that stress path will be determined, amongst other factors, by the
irreversibility and hysteresis can be expected for several reservoir stress path coefficient during injection.
reasons. So, what has been achieved is maybe more bits We have not planned to include any specific field case in
and pieces of a map that the user will need in order to this paper. Rather, the intention was to present generic
reach the final target. considerations that should be relevant for different
Material behavior is irreversible, as demonstrated by applications. For instance, the percentage change of the
laboratory experiments with synthetic rocks formed fracture permeability as a function of stress depends
under stress. A fundamental source of that irreversibility largely on the type of the rock, e.g. these properties
is the non-elastic (plastic) behavior of rocks when they would be completely different in a crystalline rock, such
are brought away from their initial virgin state. The as granite, and a sedimentary rock, such as shale or
stress path hysteresis is controlled by the product sandstone. Again, the objective here was not to perform
(1 K 0 ) , and only in the initial state and at the turning a review of available experimental databases but rather
to take a step further and to illuminate the effects that
point of the stress path is this given by the purely elastic
these properties can have in reservoir geomechanics.
rock properties. This may provide guidance towards
estimating stress path from seismic data, but then
neglecting the effects of thermal stresses caused by
injection of cold fluids, and the other mechanisms that
were considered in this paper.
Figure 8. a)The stress path coefficient h along a horizontal line Y=0 crossing the reservoir center, b) Mohr circle at the reservoir
center, c) Mohr circles at the inner flanks, d) Mohr circles at the outer flanks for different scenarios. Arrows to the right and to left
are for depleted and injected scenarios respectively. (The legend description can be found in the text and in Figure 7).

Figure 9. a) The stress path coefficient h along a horizontal line crossing the reservoir center, b) Mohr circle at the reservoir
center, c) Mohr circles at the inner flanks, d) Mohr circles at the outer flanks for only injection and injection after depletion.
of coupled properties for a single rock joint. Int. J. Rock Mech.
Min. Sci. 36: 641-650.
Fjr, E., R.M. Holt, P. Horsrud, A.M. Raaen, and R. Risnes,
2008. Petroleum Related Rock Mechanics (2nd ed.). Elsevier.
491 pp.
Fjr, E., A. Stroisz, R.M. Holt. 2013. Elastic dispersion
derived from a combination of static and dynamic
measurements. Rock Mech. Rock Eng. 46: 611-618.
Hettema, M.H.H., P.M.T.M. Schutjens, B.J.M. Verboom, and
H.J. Gussinklo, 2000. Production-induced compaction of a
sandstone reservoir: The strong influence of stress path. SPE
Reservoir Evaluation Eng. 3: 342-347.
Hettema, M.H.H., B. Bostrm, and T. Lund. 2004. Analysis of
lost circulation during drilling in cooled formations. SPE
90442; 12 pp.
Figure 10. Number of damaged elements vs pressure change Holt, R.M. and C.J. Kenter. 1992. Laboratory simulation of
and the corresponding h due to depletion and rebound. core damage induced by stress release. In Tillerson &
Waversik (eds.) Proc. 33rd US Rock Mechanics Symposium,
A.A. Balkema.
ACKNOWLEDGEMENTS
Holt, R.M., M. Brignoli, and C.J. Kenter. 2000. Core Quality:
The authors wish to thank colleagues at NTNU and Quantification of coring induced rock alteration. Int. J. Rock
SINTEF for valuable discussions. SINTEF Petroleum Mech. & Min. Sci. 37: 889-907.
Research is acknowledged (AL) for permission to Holt, R.M. and J. F. Stenebrten. 2013. Controlled laboratory
perform and publish part of the present work. The work experiments to assess the geomechanical influence of
is partly performed (RMH) within the ROSE Program at subsurface injection and depletion processes on 4D seismic
NTNU, funded by The Research Council of Norway and responses. Geophysical Prospecting 61: 476-488.
a number of industrial partners. This publication has also
Holt, R.M., E. Fjr, and A. Bauer. 2013. Static and dynamic
been produced with partial support from the BIGCCS moduli so equal, and yet so different. ARMA 13-521; 8 pp.
Centre (for SG), performed under the Norwegian
research program Centres for Environment-friendly Holt, R.M., J.F. Stenebrten, and M. Brignoli. 2014. Effects of
Energy Research (FME). The authors acknowledge the cementation on in situ and core compaction of soft sandstone.
following partners for their contributions: Gassco, Shell, ARMA 14-7397, 9 pp.
Statoil, TOTAL, ENGIE, and the Research Council of Gutierrez, M., L.E. ino, and R. Nygrd. 2000. Stress-
Norway (193816/S60). dependent permeability of a de-mineralised fracture in shale.
Mar. Petrol. Geol. 17: 895-907.
Lavrov, A. 2003. The Kaiser effect in rocks: principles and
REFERENCES stress estimation techniques. Int. J. Rock Mech. Min. Sci. 40:
Addis, M.A., N.C. Last, and N.A. Yassir. 1996. Estimation of 151-171.
horizontal stresses at depth in faulted regions and their Lavrov, A. 2005. Fracture-induced physical phenomena and
relationship to pore pressure variations. SPE Formation memory effects in rocks: a review. Strain 41: 135-149.
Evaluation, 11(1) 11-18.
Lavrov, A., I. Larsen and A. Bauer. 2015. Numerical
Alassi, H.T. 2008. Modelling reservoir geomechanics using modelling of extended leak-off test with a pre-existing
discrete element method: Application to reservoir monitoring. fracture. Rock Mechanics and Rock Engineering. 49: 1359-
PhD thesis at NTNU, Trondheim, Norway. 1368.
Becker, D., B. Cailleau, T. Dahm, S. Shapiro, and D. Kaiser. Lavrov, A. 2016a. From fracture gradient to the spectrum of
2010. Stress triggering and stress memory observed from lost-circulation pressures: a paradigm shift? Abstract
acoustic emission records in a salt mine. Geoph. J. Int. 182: submitted to GHGT-13.
933-948.
Lavrov, A. 2016b. Lost Circulation: Mechanisms and
Chan, A.W., P.N. Hagin, and M.D. Zoback. 2004. Solutions. 1st ed. Oxford. Elsevier.
Viscoplastic deformation, stress and strain paths in
unconsolidated reservoir sands (part 2): Field applications Lavrov, A. 2016c. Dynamics of stresses and fractures in
using dynamic DARS analysis. ARMA/NARMS 04-568; 9 pp. reservoir and cap rock under production and depletion. Energy
Procedia 86: 381-390.
Esaki, T., S. Du, Y. Mitani, K. Ikusada, and L. Jing. 1999.
Development of a shear-flow test apparatus and determination Longuemare, P., M. Mainguy, P. Lemonnier, A. Onaisi, Ch.
Grard, and N. Koutsabeloulis. Geomechanics in reservoir
simulation: overview of coupling methods and field case
study. Oil & Gas Science and Technology Rev. IFP 57 (5):
471-483.
Lorenz, J.C. 1999. Stress-sensitive reservoirs. J. Pet. Tech.
51(1): 61-63.
Mahi, A. (2003). Stress development in oil reservoirs. Ms.C.
thesis, NTNU. Trondheim, Norway.
Mulders, F.M.M. 2003. Modelling of stress development and
fault slip in and around a producing gas reservoir. PhD
Thesis, TU Delft, Netherlands.
Perkins, T.K. and J.A. Gonzalez. 1984. Changes in Earth
Stresses around a wellbore caused by radially symmetrical
pressure and temperature gradients. SPEJ, April 1984, 129-
140.
Santarelli, F.J., J.T. Tronvoll, M. Svennekjaer, H. Skeie, R.
Henriksen, and R.K. Bratli. 1998. Reservoir stress path: the
depletion and the rebound. SPE/ISRM 47350 in Proc.
SPE/ISRM Eurock'98 Symposium held in Trondheim, Norway,
July 8-10, 1998; Vol 2, 203-209.
Santarelli, F.J., O. Havmller, and M. Naumann. 2008.
Geomechanical aspects of 15 years water injection on a field
complex: An analysis of the past to plan the future. SPE
112944; 13 pp.
Shahri, M.P. and S.Z. Miska. 2014. Experimental
investigation of reservoir stress path hysteresis during
production-injection periods. Paper OMAE2914-
24653presented at ASME 2014 33rd Int. Conf. on Ocean,
Offshore and Arctic Engineering, San Francisco, USA, 8-13
June 2014; 8 pp.
Smart, B.G.D., J.M. Somerville, K. Edlman, and C. Jones.
2001. Stress sensitivity of fractured reservoirs. J. Pet. Sci.
Eng. 29: 29-37.
Sreide, O.K., S. Hansen, and J.F. Stenebrten. 2014.
Estimation of reservoir stress effects due to injection of cold
fluids: an example from NCS. ARMA 12-7394; 7 pp.
Zoback, M.D., and J.C. Zinke. 2002. Production-induced
normal faultingin the Valhall and Ekofisk oil fields. Pure Appl
Geophys. 159: 403-420.

Potrebbero piacerti anche