Sei sulla pagina 1di 13

International Journal of Greenhouse Gas Control 4 (2010) 827839

Contents lists available at ScienceDirect


International Journal of Greenhouse Gas Control
j our nal homepage: www. el sevi er . com/ l ocat e/ i j ggc
Geomechanical analysis of the Naylor Field, Otway Basin, Australia:
Implications for CO
2
injection and storage
Sandrine Vidal-Gilbert
a,b,
, Eric Tenthorey
a,c
, Dave Dewhurst
a,d
, Jonathan Ennis-King
a,e
,
Peter Van Ruth
b,f
, Richard Hillis
a,b
a
The Cooperative Research Centre for Greenhouse Gas Technologies (CO2CRC), Canberra, Australia
b
Australian School of Petroleum, University of Adelaide, Australia
c
Geoscience Australia, Canberra, Australia
d
CSIRO Petroleum, Perth, Australia
e
CSIRO Petroleum, Melbourne, Australia
f
Woodside, Perth, Australia
a r t i c l e i n f o
Article history:
Received 25 August 2009
Received in revised form 1 June 2010
Accepted 4 June 2010
Available online 6 July 2010
Keywords:
Otway Basin Australia
In situ stress
Reservoir stress path
Fault stability
a b s t r a c t
A geomechanical assessment of the Naylor Field, Otway Basin, Australia has been undertaken to inves-
tigate the possible geomechanical effects of CO
2
injection and storage. The study aims to evaluate the
geomechanical behaviour of the caprock/reservoir system and to estimate the risk of fault reactivation.
The stress regime in the onshore Victorian Otway Basin is inferred to be strikeslip if the maximumhori-
zontal stress is calculated using frictional limits and DITF (drilling induced tensile fracture) occurrence, or
normal if maximumhorizontal stress is based on analysis of dipole sonic log data. The NWSE maximum
horizontal stress orientation (142

N) determined from a resistivity image log is broadly consistent with


previous estimates and conrms a NWSE maximum horizontal stress orientation for the Otway Basin.
An analytical geomechanical solution is used to describe stress changes in the subsurface of the Naylor
Field. The computed reservoir stress path for the Naylor Field is then incorporated into fault reactivation
analysis to estimate the minimum pore pressure increase required to cause fault reactivation (^P
p
).
The highest reactivation propensity (for critically-oriented faults) ranges froman estimated pore pres-
sure increase (^P
p
) of 1MPa to 15.7MPa (estimated pore pressure of 18.533.2MPa) depending on
assumptions made about maximum horizontal stress magnitude, fault strength, reservoir stress path
and Biots coefcient. The critical pore pressure changes for known faults at Naylor Field range from an
estimated pore pressure increase (^P
p
) of 2MPa to 17MPa (estimated pore pressure of 19.534.5MPa).
2010 Elsevier Ltd. All rights reserved.
1. Introduction
The geological storage of carbon dioxide (CO
2
) has been pro-
posed as a potential method of reducing greenhouse gas emissions.
The Naylor Field in the Otway Basin, Victoria, has been chosen as
a demonstration site (The Otway Project) for the geological stor-
age of CO
2
by the Cooperative Research Centre for Greenhouse Gas
Technologies (CO2CRC). The Naylor Field is a small depleted nat-
ural gas eld, with the original gas cap area estimated at 40hA.
The composition of the original gas (in mole %) is 88% methane, 4%
ethane, 2% propane, 1% carbon dioxide, 2% nitrogen, and 3% other
components. Total production from the target reservoir from June
2002 to October 2003 was 9.510
7
m
3
(at standard conditions of

Corresponding author at: Total, Gas & Power, Research and Development,
CO2 Geological Storage, Paris La Defense, France. Tel.: +33 1 47 44 24 61.
E-mail address: sandrine.vidal-gilbert@total.com (S. Vidal-Gilbert).
15

C and 0.101325MPa). This was about 60% of the estimated gas


in place. Using equivalent volumes at reservoir conditions would
indicate a CO
2
storage capacity of about 210,000 tonnes, but hys-
teretic effects in relative permeability and the inux of formation
water from the adjoining aquifer may reduce this amount.
CO
2
-rich gas has been produced from the nearby Buttress Field
and injected into the CRC-1 borehole within the Naylor Structure
to demonstrate the viability of geological sequestration of CO
2
in
Australia (Fig. 1). The injected gas has an average composition of 77
mole% carbon dioxide, 20 mole% methane and 3 mole% other gas
components. Between March 18, 2008 and August 28, 2009, 65,445
tonnes of this gas wereinjectedintotheNaylor Fields WaarreCFor-
mation, containing about 58,000 tonnes of CO
2
. The reservoir was
monitored before, during and after injection via downhole pres-
sure and temperature gauges in the injection well, uid sampling
fromthe reservoir at the Naylor-1 observation well (via three level
U-tube assembly), and various geophysical methods including 4D
seismic and microseismic.
1750-5836/$ see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijggc.2010.06.001
828 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
Fig. 1. Study area location map: CO2CRC Otway Project.
Subsurface injection of CO
2
induces pore pressure variations
that affect the in situ stress state within the reservoir and its
surroundings. Injection of CO
2
has the potential to increase pore
pressure and reduce fault stability in zones within and surround-
ing the CO
2
plume. The likelihood of fault reactivation is increased
when the pore pressure becomes elevated beyond a critical pres-
sure change, which is controlled by parameters such as fault
orientation, and friction and cohesion within the fault plane. It is
thereforedesirabletoavoidexceedingthecalculatedpressurelimit,
as fault reactivation may result in cap rock failure or permeabil-
ity increase of the fault zone, both of which may result in vertical
leakage of CO
2
.
This paper outlines some of the geomechanical implications of
injecting CO
2
, with an emphasis on understanding fault stability
issues. In doing so, we also present calculations designed to incor-
porate the effects of pore pressure/stress coupling, also known as
the reservoir stress path. The reservoir stress path refers to changes
in the horizontal stress eld that are driven by variations in the
uid pressure, and is a product of complex poroelastic interac-
tions. Rather, than using a simplied effective stress law, stress
path prediction allows the evaluation of the mechanical stabil-
ity of both reservoir/caprock system and bounding faults under
the injection loading condition. This paper will present results of
fault stability modelling for the Naylor Field by incorporating stress
path modelling and considering various fault property scenarios
andconsidering several different possibilities for the contemporary
stress eld at Naylor.
2. Study area
The Naylor Field is located in the Port Campbell region, within
the onshore Otway Basin, Victoria, Australia (Fig. 1). The Nay-
lor Field is a fault-bound trap formed during the development
of the passive margin of southeastern Australia. Previous stud-
ies show that the in situ stress eld in the Otway Basin has
evolved from the normal fault stress regime associated with pas-
sive margin development during the Jurassic to Cretaceous eras
to a strikeslip or reverse regime (Jones et al., 2000; Lyon et al.,
2005; Nelson et al., 2006; Rogers et al., 2008) associated with the
Miocene-Recent compression (Schneider et al., 2004; Hillis et al.,
2008).
The target horizon for CO
2
injection is the Late Cretaceous
Waarre Formation (Figs. 13). The Waarre Formation (2002 mSS
at CRC-1 and 1977 mSS at Naylor-1) is overlain by the Flaxmans
Formation (1972 mSS at CRC-1 and 1954 mSS at Naylor-1) and
the Belfast mudstone. The Late Cretaceous Waarre Formation in
the Naylor Field originally held a methane-rich gas accumulation,
which was produced from June 2002 to October 2003. There was a
residual gas cap and the pressure increased following production
due to aquifer recharge. The pressure response to both depletion
and injection is shown for the Naylor-1 well location in Fig. 4. This
site was selectedas the locationfor a CO
2
injectionpilot project due
tothe goodporosityandpermeabilityof its reservoir rock(the aver-
age permeability was more than 1D). Furthermore, the reservoir
is overlain by the laterally extensive and thick Belfast mudstone,
which based on laboratory analyses should be able to support a
CO
2
column height in the range of 607851m (contact angle: 0

)
withanaverage of 754m(Daniel, 2007). The Naylor site is alsoclose
to the Buttress Field, a source of CO
2
-rich gas (Watson and Gibson-
Poole, 2005). The occurrence of natural high CO
2
accumulations in
the Port Campbell Embayment demonstrates that traps in the area
are capable of containing CO
2
over geological timescales (510
3
to 210
6
years: Watson et al., 2004).
There are three wells in the Naylor Field (Fig. 1). Naylor-1 was
drilled in May 2001 and discovered a natural gas accumulation in
the Waarre C Formation. Naylor South-1 was drilled in December
2001 and CRC-1 was drilled by the CO2CRC in March 2007. For
the Otway Project, CRC-1 was used as the CO
2
injection well, with
Naylor-1 being the updip monitoring well.
The Naylor Field is bound to the west by a northsouth trend-
ing normal fault (Naylor Fault). The Naylor Fault has an effective
juxtaposition seal because fault throwis insufcient to completely
offset the seal (Belfast mudstone). The Naylor Fault forms part of
the structural closure which contains the injected CO
2
plume, and
is required to act as a long-term seal. The Naylor structure is also
cut to the east by a normal fault (Naylor East Fault) and it is bound
to the South by the Naylor South Fault (Fig. 3). Neither the Naylor
East Fault nor the Naylor South Fault is in the expected migration
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 829
Fig. 2. CRC-1 and Naylor-1 well section with formation tops (depths are in mSSTVD).
pathway of the injected CO
2
plume. The faults bounding the Nay-
lor eld supported the initial natural gas column (initial Gas Water
Contact was 2015mSS; Spencer et al., 2006), and the injected vol-
ume of CO
2
at subsurface conditions was smaller than the volume
of produced methane at the same conditions. Therefore, the faults
bounding the Naylor Field should have sufcient sealing capacity
to hold the CO
2
volume injected. However, more research has to
be conducted regarding the sealing capacity of faults and howthey
will respond to the different wettability and density of CO
2
and
CH
4
. In this paper, the analysis will be focussed on potential fault
Fig. 3. Composite from the 3D seismic reection survey going through the key wells with interpretated seismic horizons and main faults (Courtesy of T. Dance, CO2CRC).
830 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
reactivation, which may lead to increased fault zone permeability,
as a potential containment risk for CO
2
storage projects within the
Waarre Formation in the Naylor Field.
3. Geomechanical model
The strength of a rock, and how much stress the rock supports
must be well constrained when trying to determine the reser-
voir/caprockintegrityandfault stability. The geomechanical model
consists of in situ stress and rock strength data and provides the
basis for all geomechanical studies. Thegeomechanical model of the
onshore Victorian Otway Basin is outlined in the following section.
3.1. Neotectonic records and stress indicators from earthquakes
Estimation of the in situ stress state from petroleum data
combinedwithearthquake focal mechanismsolutions andthe neo-
tectonic record provide important insights into the structural and
tectonic history of the region. Nelson et al. (2006) have discussed
the insitustress state of southeast Australia andcomparedthis with
earthquake focal mechanismsolutions and the neotectonic record.
Their overviewprovides valuable baseline data for this geomechan-
ical study.
Focal mechanismsolutions reveal stresses inthe deeper seismo-
genic zone, whichinSE Australia are typically between5 and 20km
(Allenet al., 2005). Comparingtheinsitustresses fromwell dataand
stress indicators from earthquakes allows investigation into stress
differences betweens basins and underlying basement. The Otway
Basin is relatively aseismic but dominantly strikeslip focal mech-
anisms have been recorded to the north at Nhill in Victoria (east
of Victorian border, Fig. 1). The late-Neogene to recent geological
records of SE Australia indicate signicant periods of faulting and
deformation (Dickinson et al., 2002; Sandiford et al., 2004), with
evidence for reverse faulting in the neotectonic record close to the
Victorian Otway Basin (Otway Ranges, Minerva anticline, Fig. 1).
3.2. In situ stress assessment
The geomechanical integrity of the reservoir is controlled by the
stress regime at the site and by the injection pressure. Stresses are
tensorial in nature and are characterized by magnitudes and orien-
tations of the three principal components magnitudes, o
1
, o
2
and
o
3
which are orthogonal. A basic assumption is that the principal
stress directions are approximately vertical and horizontal. In this
case, the principal stresses are denoted o
V
for the vertical stress,
and o
Hmax
and o
hmin
for the maximum and minimum horizontal
stresses, respectively.
3.2.1. Orientation of maximum horizontal stress
The most commonly used information for inferring stress ori-
entations is given by borehole breakouts. These symmetric spalled
regions are formed at various depths on the wellbore wall where
the compressive stress concentration exceeds the shear strength
of the rock. In vertical wells through transversely isotropic rocks,
breakout elongates the wellbore parallel to o
hmin
(Zoback, 2008).
The orientation of maximum horizontal stress was determined to
be N1425

E from breakouts in the CRC-1 borehole interpreted


on a resistivity image log (FMI) (Van Ruth, 2007). This maximum
horizontal stress orientationis broadly consistent withthe regional
orientation and conrms a NWSE maximumhorizontal stress ori-
entation in the onshore Victorian Otway Basin (Hillis et al., 1998;
www.asp.adelaide.edu.au/asm).
3.2.2. Magnitude of vertical stress
Vertical stress (o
V
) is the stress applied at any given point due
to the weight of the overlying rock mass and uids. Vertical stress
Fig. 4. Measured pore pressure at CRC-1 and Naylor-1 wells.
magnitude can be estimated by integrating the bulk density log of
the overlying uid-saturated rock with depth:
o
V
=
_
D
0

b
gdz (1)
where g is the gravitational acceleration (9.81m/s
2
), Dis depth and

b
is the bulk density of the uid-saturated rock.
Density logs used for estimating o
V
should be as complete as
possible. For the upper unlogged interval, bulk density was esti-
mated using check-shot log velocities with empirical relationships
linking velocities and densities. The density predictions are made
using the lithology-specic polynomial forms of the Gardner et al.
(1974) velocitydensity relationships improved by Castagna et al.
(1993). Vertical stress values obtained for the onshore Victorian
Otway Basin indicate an average vertical stress gradient of about
21.45MPa/km.
3.2.3. Magnitude of minimum horizontal stress
The minimum horizontal stress can be estimated by various
means. One way is using micro- and mini-fracture tests, extended
leak-off tests and massive hydraulic fracture records to interpret
the fracture closure pressure, which corresponds to the minimum
horizontal stress magnitude. Conventional leak-off tests are com-
pleted once leak-off occurs and as such, it is not possible to record
fracture propagation, shut-in response and fracture closure. Con-
sequently, if several leak-off pressure data are available, a lower
bound to these data should provide a reasonable estimate of the
minimum horizontal stress magnitude (Hawkes et al., 2005).
Nelson et al. (2006) have gathered a series of leak-off pres-
sures, recorded in well completion reports and a series of
leak-off tests, performed in wells across the Victorian Otway Basin
(Fig. 5a). The average of these measurements indicates a gradi-
ent of 18.5MPa/km. The lower bound to these data is around
15.5MPa/km. Inaddition, anextendedleak-off test was undertaken
withinthe 512519mKBdepthinterval during the drilling of CRC-1
(VanRuth, 2007). The gradient determinedfromthe extendedleak-
off test of CRC1 well is 14.5MPa/km. This test has been performed
at relatively shallow depth compared to the target reservoir depth
(2000mSS).
Berard et al. (2008) used extended leak-off test data, a bore-
hole wall electrical image and dipole sonic log data from the well
CRC-1 to constrain the principal horizontal stress orientations and
magnitudes. They conclude that the minimum and maximum hor-
izontal stress gradients are on average, equal to 16MPa/km and
18MPa/km, respectively.
As no tests were undertaken at reservoir depth, the minimum
horizontal stress is poorly constrained. To consider all potential
assumptions, minimumhorizontal stress gradients of 14.5MPa/km
and of 18.5MPa/km are used here.
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 831
Fig. 5. (a) Minimum horizontal stress estimates in the onshore Victorian Otway Basin and (b) Polygons which dene possible stress magnitudes for Naylor eld. Red line
shows the possible range for oHmax assuming o
hmin
=14.5MPa/km(fromCRC-1 ELOT). (For interpretation of the references to colour in this gure legend, the reader is referred
to the web version of the article.)
3.2.4. Magnitude of maximum horizontal stress
The Frictional Limit method and the occurrence of DITF (Drilling
Induced Tensile Fractures) observed in two of the four image logs
in the Victorian Otway Basin (Nelson et al., 2006) have been used
to estimate the maximumhorizontal stress. Frictional limits theory
states that the ratio of the maximum to minimum effective stress
cannot exceed the magnitude required to cause faulting on a criti-
cally oriented, pre-existing, cohesionless fault plane (Sibson, 1974).
Thus the magnitude of the maximum horizontal stress can be con-
strained when the magnitude of the minimum horizontal stress is
known (Moos and Zoback, 1990). The frictional limit to stress is
given by:
o
1
P
p
o
3
P
p

_
_
(
2
+1) +
_
2
(2)
where is the coefcient of friction, P
p
is the pore pressure, is
the Biots coefcient, o
1
is the maximum principal stress and o
3
is
the minimum principal stress.
Pore pressure was measured using Schlumbergers Modular
Dynamics Tester (MDT) tool in the CRC-1 borehole before and dur-
ing CO
2
injection (Fig. 4). Fig. 4 gives also pressure variations at
Naylor-1 well during methane production and before CO
2
injec-
tion. The maximumprincipal stress is assumed to be the maximum
horizontal stress. Using Eq. (2), the maximum value for the max-
imum horizontal stress gradient in the onshore Victorian Otway
Basin was constrained to 27MPa/km at reservoir level using the
in situ stress gradient determined herein (o
hmin
14.5MPa/km,
o
V
21.45MPa/km, P
p
8.64MPa/km just before CO
2
injection,
February 2008) and assuming a coefcient of friction =0.6.
Nelson et al. (2006) used the occurrence of DITFs, knowl-
edge of the o
hmin
and o
V
gradient (o
hmin
18.5MPa/km,
o
V
21.45MPa/km) andthe assumptionthat the tensile strengthof
the reservoir rocks are negligible to constrain the o
Hmax
gradient to
about 37MPa/kmintheVictorianOtwayBasin. Takingtheextended
leak-off test measurement from CRC-1 (o
hmin
14.5MPa/km), the
occurrence of DITFs allow us to constrain the o
Hmax
gradient to
about 26MPa/km in the Onshore Victorian Otway Basin.
Fig. 5b illustrates the allowable stress states at a depth of
2025mGL (1977mSS, the Waarre C top formation at Naylor-1),
assuming that stress accumulation is limited by frictional limit
theory. The red line shows the range of possible values for the
maximumhorizontal stress gradient, using a value of 14.5MPa/km
for the minimum horizontal stress gradient. Depending on the
value used for the maximum horizontal stress, the faulting regime
may be either normal or strikeslip. DITF data and Frictional Limit
theory presented below suggest a strikeslip faulting regime and
the inversion of sonic scanner data from CRC-1 well results in a
normal fault regime. In this study, three scenarios with different
assumptions for the stress regime (strikeslip fault regime: SSFR
and normal fault regime: NFR) given in Table 1, have been used in
a later section for assessing the risk of fault/fracture reactivation.
This described stress state is considered for the further modelling
as the initial stress state after depletion and just before CO
2
injec-
tion, inFebruary2008. Inadditiontothepreviouslydescribedinsitu
stress regime, the stress alteration induced by CO
2
injection has to
be determined for a better estimate of the fault reactivation risk.
This stress alteration is often identied as the reservoir stress path.
4. Reservoir stress path
The stresses acting within a reservoir are characterized by three
orthogonal stresses which are approximately vertical and horizon-
tal. The two horizontal stresses are a combination of the lateral
effect of the overburden, the Poisson effect, plus any tectonic stress
change, or geometric constraint which results in different hori-
zontal stress magnitudes (Addis, 1997). The pore pressure within
the formation also affects the horizontal stress magnitudes, both
in the initial state and during production. Exploitation of under-
ground resources causes perturbation to the pore pressure prole.
If pore pressure changes during production/injection, the evolution
of the stresses withproduction/injectionshouldalsobe considered,
as stress and pore pressure magnitudes are intrinsically linked. In
recent years there has beenincreasing evidence fromoil eldreser-
voirs that changes to pore pressure may also impact directly on
the regional stress magnitudes due to complex poroelastic effects
832 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
Table 1
In situ stress tensor for a strikeslip fault and normal fault regime assumptions.
Scenario oV gradient
(MPa/km)
o
hmin
gradient
(MPa/km)
oHmax gradient
(MPa/km)
Pp gradient
(MPa/km)
oHmax orient.
(N)
Scenario 1: SSFR 21.45 14.5 26 8.64 142
Scenario 2: NFR 21.45 14.5 18 8.64 142
Scenario 3: SSFR 21.45 18.5 37 8.64 142
Fig. 6. Mohr circles, failure envelope and variation of Poissons ratio with effective conning pressure from laboratory tests on sandstones from the Waarre C Formation at
2056.4mKB.
(Segall, 1989; Grasso, 1992; Addis, 1997; Hillis, 2001). This effect
is known as the reservoir stress path or stress-depletion response
(Addis, 1997) or pore pressure-stress coupling (Hillis, 2001) and
is referred to as a decrease/increase in the minimum horizontal
stress accompanying depletion/injection. The reservoir stress path
is dened as the ratio of the change in minimum horizontal stress
(o
hmin
) to the change in pore pressure (P
p
), and usually has a value
of 0.50.8 (Addis, 1997). Unfortunately, despite numerous obser-
vations, this phenomenon remains rather poorly understood. The
reservoir stress path is not known before exploitation (produc-
tion and/or injection) and analytical or numerical models for stress
development in reservoirs are very sensitive to the input param-
eters. Furthermore, in some cases in the North Sea, some sort of
irreversibility has been observed in terms of reservoir path upon
re-pressurisation. The reservoir did not followthe same stress path
during depletion and during pressure rebound (Santarelli et al.,
1998).
Understanding the reservoir stress path during both depletion
and re-pressurisation is important for estimating the reservoir
compaction/expansion, surface movement, failure of intact rock
and near wellbore deformation, and it is required for identifying
minimum pore pressure required to cause fault reactivation. The
ideal procedure is to measure o
Hmax
and o
hmin
with in situ stress
measurements at initial reservoir conditions and at one or more
stages of pore pressure changes (Rhett and Risnes, 2002). Lacking
repeated in situ stress measurements, some analytical models, e.g.
uniaxial strain conditions and Eshelbys solution (Rudnicki, 1999),
are used in this paper to estimate reservoir stress path.
4.1. The approach to uniaxial strain
It follows fromlinear poroelasticity that a reservoir will behave
under uniaxial strain conditions such that the reservoir stress path
equals:
=
^o
h
^P
p
=
_
1 2
1
_
(3)
where is the Biots coefcient (or effective stress parameter) is
usually assumed to be 1, but for sandstone this is not always the
case (Bouteca, 1994; Hettema et al., 1998; Addis, 1997). As Biots
coefcient is not always 1for sandstones, a sensitivity analysis with
=0.7 and =1 is performed here. Triaxial testing was undertaken
on core samples from Waarre Formation Unit C in the CRC-1 bore-
hole. The failure envelope shows a cohesive strength of just above
5MPa and a friction coefcient of 0.76 (Fig. 6). The Poissons ratios
at different effective conning pressures for the sandstone reser-
voir rock are given in Fig. 6. The Poissons ratio ranges from 0.22
to 0.32 so the resulting pore pressure-stress coupling ratio ranges
from =0.37 to =0.71, assuming =0.7 and =1.
4.2. The solution of Eshelby
Rudnicki (1999) extended the solution of Eshelby (1957) to cal-
culate the effects of geometry and elastic properties on altering the
local stress state. In this model, the theory of inhomogeneities is
used to solve induced stress changes within an ellipsoidal reservoir
(inhomogeneity) embedded in a surrounding material (host rock)
with different elastic properties. The formulations for injection or
withdrawal of uid from a reservoir given in Rudnicki (1999) are
used in this paper. In the following equations, the subscripts I and
stand for inhomogeneity (reservoir) and surrounding material,
respectively. The principal semi-axes of the ellipsoid are a and c
(with a =b in the horizontal plane) and the aspect ratio of the inho-
mogeneity is e =c/a. Rudnicki demonstrated that the lateral stress
increment is:
^o
h
=
I
zP
I
p
_
1
3
(1 +2R) +
3
(1 R)
(1 2
I
)
(1 +
I
)
_
(4)
where
I
is the Biots coefcient of the inhomogeneityandthe other
terms are dened below.
For an axisymmetric reservoir, the ratio of lateral to axial strain
increments is given by the following expression:
R =
3u S
33kk
+g
_
3uS
3333
S
33kk
S
pp33
_
2S
33kk
+g
_
3uS
3333
S
33kk
S
pp33
_ (5)
With:
S
33
= 1
(1 2

)
2(1

)
l(c)
c
2
(2 3l(c))
2(1

)(1 c
2
)
, S
kk33
= 1
(1 2

)
(1

)
l(c)
S
33kk
=
(1 +

)
(1

)
(1 l(c)), l(c) =
c
2
(1 c
2
)
3]2
_
ur cos(c) c(1 c
2
)
1]2
_
u=
1
3
(1 +

)
(1

)
, g =
C
I
C

1,
3
=
u(1 k)
(1 +2R)(1 +uk) +g(1 R)(S
kk33
u)
and k =
K
I
K

1
(6)
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 833
Fig. 7. Reservoir stress path against inhomogeneity Poissons ratio (a) for different ratios of bulk moduli and Biots coefcient of 1, (b) for different ratios of shear moduli and
Biots coefcient of 1, (c) for different ratios of bulk moduli and Biots coefcient of 0.7 and (d) for different ratios of shear moduli and Biots coefcient of 0.7.
Based on the laboratory data from Fig. 5, the Poissons ratio of
the inhomogenity (
I
) ranges from 0.22 to 0.32; Rudnicki (1999)
assumed that the dependence on Poissons ratio of the surround-
ing material (

) is weak. G
I
/G

and K
I
/K

are inferred from sonic


log data. P-wave and S-wave velocities and density logs are used
to compute dynamic undrained moduli. Using Biot-Gassmanns
equation, a saturation correction is applied to dynamic undrained
moduli to obtain dynamic drained moduli. Then, empirical rela-
tionships between drained static and drained dynamic Youngs
moduli are applied (Wang, 2000; Vidal-Gilbert et al., 2009). This
approach gives an approximate value for G
I
/G

=2 and K
I
/K

=2.
The aspect ratio of the inhomogeneity (e) used in this evaluation is
0.0187. As the actual reservoir geometry is not fully represented
by an axisymetric ellipsoid, a sensitivity analysis has been car-
ried out with different values ranging from 0.018 to 0.03 for the
aspect ratio in order to evaluate the impact on the stress path
evaluation. The results show that this parameter does not have a
major inuence on the stress path estimation for this particular
reservoir.
Fig. 7a and c presents the reservoir stress path (ratio of lateral
stress increment to pore pressure increment) against the Poissons
ratioof theinhomogeneityfor different ratios of bulkmoduli andfor
a Biots coefcient of 0.7 and 1, respectively. Fig. 7b and d presents
the reservoir stress path (ratio of lateral stress increment to pore
pressure increment) against the Poissons ratio of the inhomogene-
ity for different ratios of shear moduli and for a Biots coefcient
of 0.7 and 1, respectively. Assuming that the Poissons ratio of
the inhomogenity ranges from 0.22 to 0.32, Fig. 7ad shows that
the reservoir stress path ranges from =0.36 to =0.75 assuming
G
I
/G

=2, K
I
/K

=2 and =0.7 and 1. The results of this solutionare


used for the reservoir stress path estimation (=0.4 and =0.8) to
study the likelihood of fault reactivation during CO
2
injection.
5. Geomechanical risking
The injectionof CO
2
into the subsurface will result inanincrease
in the reservoir pore pressure. Increasing pore pressure can cause
brittle failure of rocks, which will occur when the stress acting on
a rock exceeds rock strength. The maximum pore pressure which
can be sustained by faults and intact rock can be estimated from
geomechanical risking (Root et al., 2004; Streit and Hillis, 2004).
Inthis paper, the reservoir stress pathandfault stability analysis
werenot studiedduringthedepletionphase. Thepresentedinitial
state is considered after depletion and before CO
2
injection.
5.1. Pore pressure required for inducing faulting
Inducing slip on an inactive fault provides a possible path for
leakage. Slip will occur on a fault when the maximum shear stress
acting in the fault plane exceeds the shear strength of the fault. In
a 2D analysis, the magnitudes of the shear stress (z) and normal
stress (o
n
) acting on this plane are given by:
z =
o
1
o
3
2
sin 20 and o
n
=
o
1
+o
3
2
+
o
1
o
3
2
cos 20 (7)
834 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
whereo
1
is themaximumprincipal insitustress, o
3
is theminimum
principal insitustress and0 is the angle betweenthe fault plane and
the o
3
direction. A MohrCoulomb shear failure criterion is then
used to characterize the fault strength:
z = c +
_
o
n
P
p
_
(8)
where z is the critical shear stress for slip to occur, c is the fault
cohesion, is the fault friction coefcient (=tan where is the
fault frictionangle), o
n
is thenormal stress, is theBiots coefcient
and P
p
is the pore pressure in the fault plane. The faults are often
assumed to be cohesionless and the friction coefcient is typically
in the range of =0.6 to 0.85 (Byerlee, 1978).
Substituting Eq. (7) into Eq. (8), the pore pressure required to
reactivate fault is expressed as followed:
P
p
=
1

_
1
2
(o
1
+o
3
) +
1
2
(o
1
o
3
) cos 20
1
2
(o
1
o
3
)
sin20

_
(9)
5.1.1. Normal fault regime (NFR)
In normal fault stress regimes, the maximumprincipal stress o
1
is vertical and is denoted o
V
and the minimum principal stress o
3
is horizontal and is denoted o
h
. The faults which are most likely
to slip rst in any setting are those that contain the intermediate
principal stress axis. Insuchacase, theintermediateprincipal stress
(o
2
=o
H
) can be neglected (Hawkes et al., 2004).
The reservoir stress path ratio can be combined with the
MohrCoulomb criterion for failure in normal fault stress regimes
and a newequation is derived for the pore pressure injection levels
that can induce slip on faults. The total horizontal stress (o
h
) can
be written as function of the initial pressure (P
pi
) and the change
in total horizontal stress (^o
h
) induced by injection:
o
h
= o
h0
+
^o
h
^P
p
(P
p
P
pi
) or o
h
= o
h0
+(P
p
P
pi
) (10)
The failure criterion given in Eq. (9) can nowbe rewritten using
Eq. (10):
P
p
=
1

_
(1]2)(o
v
+o
h
) +(1]2)(o
v
o
h
) cos 20
(1]2)(o
v
o
h
)(sin20])
1 (1]2)(1 cos 20 +(sin20]))

(1]2)P
pi
(1 cos 20 +(sin20]))
1 (1]2)(1 cos 20 +(sin20]))
_
(11)
For anormal fault stress regime(scenario2, Table1), 0 is thefault
dip angle. Calculations of pore pressure levels required to cause
faulting are conducted for the Naylor eld using the normal fault
in situ stress assumption given in Table 1. The total vertical stress
is approximately 43.4MPa, while the minimum horizontal stress
is 29.4MPa at a pore pressure (P
pi
) of 17.5MPa in the initial state
within the reservoir at a depth of 2025mGL (1977mSS, the Waarre
C reservoir top formation at Naylor-1).
Fig. 8shows pore pressure that is estimatedto cause fault reacti-
vation assuming that the total horizontal stress are constant (=0),
and that Biots coefcient =1. In this conguration, the increase
in pore pressure required to reactivate a fault with a dip angle of
60

is ^P
p
=5.3MPa, with P
pi
=17.5MPa. For the same pore pres-
sure increase and considering a reservoir stress path of =0.4 and
of =0.8, the stress state is far from the failure line. Regarding
=0.4 scenario, the pore pressure increase required to cause fault
reactivation is ^P
p
=12.9MPa and regarding =0.8 scenario, fault
stability is never jeopardized, even at large pore pressures. Table 2
summarizes pore pressure increase required to cause fault reacti-
vation (^P
p
) assuming a normal fault stress regime (scenario 2), a
Biots coefcient of =0.7 or 1 and a reservoir stress path of =0
or 0.4.
Fig. 8. Scenario 2: NFR MohrCoulomb circle assuming a pore pressure varia-
tion of 5.3MPa without any pore pressure/stress coupling ratio (=0), with a pore
pressure/stress coupling ratio =0.4 and =0.8, assuming =1.
5.1.2. Strikeslip fault regime (SSFR)
The pore pressure/stress coupling ratio has not been clearly
established for maximum horizontal stress for strikeslip stress
regimes. Hawkes et al. (2004) recommend using site-specic, cou-
pled reservoir-geomechanical simulations for such conditions.
Nevertheless, the maximumhorizontal stress path and the min-
imumhorizontal stress pathhavebeenassumedtobesimilar (Rhett
and Risnes, 2002) so Eq. (9) becomes:
P
p
=
1

_
(1]2)(o
H
+o
h
) +(1]2)(o
H
o
h
) cos 20
(1]2)(o
H
o
h
)(sin20])
1 (1]2)

P
pi
1
_

_
(12)
For a strikeslip fault regime, 0 is the angle between the strike
of a vertical fault and o
h
.
Calculations of pore pressure levels required to cause faulting
are conducted for the Naylor eld using the strikeslip fault in situ
stress assumptions given in Table 1.
For scenario 1 given in Table 1, the maximum horizontal stress
is 52.6MPa, while the minimum horizontal stress is 29.4MPa at a
porepressureof 17.5MPaat theinitial statewithinthereservoir at a
depthof 2025mGL(1977mSS, theWaarreCreservoir topformation
at Naylor-1).
Fig. 9 shows pore pressure that is estimated to cause fault
reactivation assuming that the total horizontal stresses are con-
stant (=0) and that Biots coefcient is 1. When the total
horizontal stresses are assumed constant, the increase in pore
Fig. 9. Scenario 3: SSFR MohrCoulomb circle assuming a pore pressure varia-
tion of 2MPa without any pore pressure/stress coupling ratio (=0), with a pore
pressure/stress coupling ratio =0.4 and =0.8, assuming =1.
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 835
Table 2
Pore pressure increase (^Pp) required to reactivate critically oriented faults depending on assumptions made about in situ stress regime, fault strength, reservoir stress path
and Biots coefcient.
Scenario Stress regime Fault strength Reservoir stress path Biots coefcient ^Pp (MPa) Pp (MPa)
Scenario 1 SSFR Cohesionless faults =0 =1 1 18.5
SSFR Cohesionless faults =0.4 =1 1.8 19.3
SSFR Cohesionless faults =0 =0.7 8.9 26.4
SSFR Cohesionless faults =0.4 =0.7 9.9 27.4
SSFR Healed faults =0 =1 10.8 28.3
SSFR Healed faults =0.4 =1 11.5 29
SSFR Healed faults =0 =0.7 22.9 40.4
SSFR Healed faults =0.4 =0.7 23.9 41.4
Scenario 2 NFR Cohesionless faults =0 =1 5.3 22.8
NFR Cohesionless faults =0.4 =1 12.9 30.4
NFR Cohesionless faults =0 =0.7 15.1 32.6
NFR Cohesionless faults =0.4 =0.7 25.9 43.4
NFR Healed faults =0 =1 13.9 31.4
NFR Healed faults =0.4 =1 20.7 38.2
NFR Healed faults =0 =0.7 27.3 44.8
NFR Healed faults =0.4 =0.7 37 54.5
Scenario 3 SSFR Cohesionless faults =0 =1 2.3 19.8
SSFR Cohesionless faults =0.4 =1 3.8 21.3
SSFR Cohesionless faults =0 =0.7 10.8 28.3
SSFR Cohesionless faults =0.4 =0.7 12.9 30.4
SSFR Healed faults =0 =1 14.3 31.8
SSFR Healed faults =0.4 =1 15.7 33.2
SSFR Healed faults =0 =0.7 27.9 45.4
SSFR Healed faults =0.4 =0.7 29.9 47.4
pressure required to reactivate a fault is ^P
p
=1MPa, with
P
pi
=17.5MPa. In contrast, the increase in pore pressure is approx-
imately ^P
p
=1.8MPa and ^P
p
=5MPa when reservoir stress path
followed by the in situ stresses during CO
2
injection is =0.4 and
=0.8, respectively. Inthis insitustress regime, the size of the circle
is not changed because it has been assumed that the minimumhor-
izontal stress path is the same as the maximum horizontal stress
path. Table 2 summarizes the pore pressure increase required to
cause fault reactivation (^P
p
) assuming a strikeslip fault stress
regime (scenarios 1 and 3), a Biots coefcient of =0.7 or 1 and a
reservoir stress path of =0 or 0.4.
5.2. Fault stability analysis
The risk of fault reactivation is calculated using the 3D formu-
lation in Eqs. (7)(12) and the geomechanical model described
in Table 1. This technique determines fault reactivation risk by
estimating the increase in pore pressure required to cause fault
reactivation (Mildren et al., 2002; Streit and Hillis, 2004). The mag-
nitude of the normal stress across the fault and the shear stress
are calculated through 3D relationships established by a change of
Cartesian reference system from the stress tensor across any fault.
The minimum pore pressure increase required to cause reac-
tivation for cohesionless faults in the Otway Basin at 2025m is
shown in Fig. 10. This gure presents plots of poles to planes,
assuming =0 and =1, for the three stress regime scenarios
described in Table 1. The orientation of faults with high and low
fault reactivation propensity differs for faults when the maximum
horizontal stress was predicted assuming a strikeslip fault regime
and when the maximumhorizontal stress was predicted assuming
a normal fault regime. In the strikeslip fault regime assumption,
sub-vertical faults that strike roughly 60

from the minimum hor-


izontal stress have the highest fault reactivation propensity (hot
colours in Fig. 10a and c). The highest reactivation risk for critically
oriented cohesionless faults is estimated to be 1MPa for scenario 1
and 2.3MPa for scenario 3. In the normal fault regime assumption,
faults that strike sub-parallel to the maximum horizontal in situ
stress and dip at roughly 60

have the highest fault reactivation


propensity (hot colours in Fig. 10b). The highest reactivation risk
for critically oriented cohesionless faults in scenario 2 is estimated
to be 5.3MPa.
The same risk of fault reactivation is presented below incorpo-
rating into equations of fault stability the estimated stress paths
followed by the in situ stresses during CO
2
injection. The fault
stability analysis incorporating reservoir stress path (=0.4) is
illustrated on Fig. 11. Incorporating the reservoir stress path ()
into the fault stability equations gives a higher value for the estima-
tion of maximumsustainable pore pressure. This tendency is more
pronounced for normal fault stress regime assumption (scenario
2) where the minimum ^P
p
is 12.8MPa assuming =0.4 whereas
^P
p
is 5.3MPa assuming =0.
A sensitivity analysis has been performed to estimate the
fault reactivation propensity depending on assumptions made
about maximum horizontal stress magnitude (scenarios 1, 2 and 3
described in Table 1), fault strength (cohesionless faults with C=0
and =0.6 and healed faults with C=5MPa and =0.76), reservoir
stress path (=0, constant horizontal stress and =0.4) and Biots
coefcient (=0.7 and =1).
Table 2summarizes the results of this analysis. The highest reac-
tivation propensity (for critically oriented faults) ranges from an
estimatedpore pressure increase (^P
p
) of 137MPa, withaninitial
pore pressure at the top of the reservoir (P
pi
) of 17.5MPa.
Among the sensitivity analysis results, the most risky scenarios
are cohesionless faults when:
1. SSFR with =1 and =0 or 0.4, with low horizontal stress sce-
nario: ^P
p
=1MPa, P
p
=18.5MPa
2. SSFR with =1 and =0 or 0.4, with high horizontal stress sce-
nario: ^P
p
=2.3MPa, P
p
=19.8MPa
3. NFR with =1 and =0: ^P
p
=5.3MPa, P
p
=22.8MPa
In addition, it could be noted that non-zero values of reservoir
stress path () and non-unity values of Biots coefcient (), both
of which are likely, decrease the risk of fault reactivation in all
scenarios.
836 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
Fig. 10. Stereonets showing the fault reactivation propensity (^Pp) at 2025m depth in the Otway Basin. Faults are plotted as poles to planes. The results are presented for
cohesionless faults assuming a reservoir stress path of =0 (constant horizontal stresses), a Biots coefcient of =1 (a) for scenario 1, strikeslip fault stress regime (b) for
scenario 2, normal fault stress regime and (c) for scenario 3, strikeslip fault stress regime.
Fault reactivation propensity has also been evaluated for three
key faults within the Naylor structure with known orienta-
tions using maximum horizontal stress calculations from Table 1
(Fig. 12). Fault reactivation propensity is calculated using the cohe-
sionless fault strength scenario. Faults are coloured according to
reactivation propensity (^P
p
). High values of ^P
p
(cool colours)
indicate low reactivation propensity, whereas low values of ^P
p
(warm colours) indicate high reactivation propensity. Comparison
of the three model runs shows that the normal fault regime results
in a more stable fault condition, in which larger increases in pore
pressure can be supported. The fault segment with highest fault
reactivation propensity in the Naylor Field is at the base of the Nay-
lor Fault near Naylor-1 well, when fault reactivation propensity is
calculated with a SSFR regime assumption (scenarios 1 and 3).
6. Discussion
The minimum pore pressure increase required to cause fault
reactivation (^P
p
) for critically oriented faults ranges from 1MPa
to 37MPa, with an initial pore pressure at the top of the reser-
voir (P
pi
) of 17.5MPa, depending on assumptions made about
stress regime, fault strength, reservoir stress paths and Biots
coefcient. Two fault strength scenarios were used to evalu-
ate the potential for fault reactivation; healed faults (C=5MPa
and =0.76) and cohesionless faults (C=0MPa and =0.6).
In addition, three stress regimes have been considered:
SSFR with o
hmin
=14.5MPa/km and o
Hmax
=26MPa/km; NFR
with o
hmin
=14.5MPa/km and o
Hmax
=18MPa/km; SSFR with
o
hmin
=18.5MPa/km and o
Hmax
=37MPa/km. The vertical stress
gradient is constant (o
V
=21.45MPa/km) for all cases. The resultant
maximum horizontal stress magnitudes suggested a strikeslip
fault regime where the occurrence of DITF was used and a normal
fault regime where the CRC-1 sonic log inversion was used.
Therefore, fault reactivation analyses differ in terms of which
fault orientations have high or low fault reactivation propensity
depending on the method that was used to calculate maximum
horizontal stress.
Taking into account the model uncertainties, a sensitivity
analysis has been performed to estimate the fault reactiva-
tion propensity for critically oriented faults. The most risky
scenarios are a cohesionless fault in a strikeslip regime
(^P
p
=1MPa, P
p
=18.5MPa, with low horizontal stress gradient
and ^P
p
=2.3MPa, P
p
=19.8MPa, with high horizontal stress gradi-
ent). The normal fault regime is a less risky scenario (^P
p
=5.3MPa,
P
p
=22.8MPa).
The risk of the fault reactivation presented incorporates stress
paths followed by the in situ stress within the reservoir during CO
2
injection with equations of fault stability. This model shows the
important role of stress path on fault stability. Non-zero values of
reservoir stress path (=0.4) means that horizontal stress is not
constant, decreasing the risk of fault reactivation in all scenarios:
^P
p
=1.8MPa, P
p
=19.5MPa for scenario 1 (SSFR); ^P
p
=12.9MPa,
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 837
Fig. 11. Stereonets showing the fault reactivation propensity (^Pp) at 2025m depth in the Otway Basin. Faults are plotted as poles to planes. The results are presented for
cohesionless faults assuming a reservoir stress path of =0.4, a Biots coefcient of =1 (a) for scenario 1, strikeslip fault stress regime (b) for scenario 2, normal fault stress
regime and (c) for scenario 3, strikeslip fault stress regime.
P
p
=30.3MPa for scenario2 (NFR); and ^P
p
=3.8MPa, P
p
=21.3MPa
for scenario 3 (SSFR), cohesionless fault with =0.4, instead
of ^P
p
=1MPa, P
p
=18.5MPa, ^P
p
=5.3MPa, P
p
=22.8MPa and
^P
p
=2.3MPa, P
p
=19.8MPa for scenarios 1, 2 and 3 respectively,
cohesionless fault with =0, at the top of the reservoir at Naylor-1
well. However, this result depends on the assumption made about
the maximumhorizontal stress path. Lacking repeated in situ stress
measurements, some analytical models were used and both min-
imum and maximum horizontal stress paths were assumed to be
equal.
Inaddition, non-unity of the Biots coefcient (=0.7) decreases
further the risk of fault reactivation. The uncertainty linked to this
parameter could be minimized with appropriate laboratory mea-
surements. This work is planned using cores froman adjacent eld
to infer the Biots coefcient value.
Fault reactivation is one of the geomechanics-related risk fac-
tors for loss of containment of injected CO
2
. Hydraulic fracturing
and especially the risks associated with out-of-reservoir fracture
growth are to be avoided. To avoid migration through new frac-
tures, the pore pressure must remain below the fracture gradient
to ensure that fracturing is not induced. A conservative upper
bound on injection pressure is the magnitude of the minimum in
situ stress (o
3
) (Hawkes et al., 2005). In this study, the minimum
in situ stress is the minimum horizontal stress for all scenarios:
o
hmin
=29.3MPa for scenarios 1 and 2 and o
hmin
=37.5MPa for sce-
nario3. This thresholdfor injectionpressureprecludes someresults
presented in light yellow in Table 2. As a result, the possible val-
ues for minimum pore pressure increase required to cause fault
reactivation (^P
p
) for critically oriented faults ranges from 1MPa
to 15.7MPa, with an initial pore pressure at the top of the reser-
voir (P
pi
) of 17.5MPa (estimated pore pressure of 18.533.2MPa),
considering the magnitude of the minimum in situ stress as the
threshold for injection pressure. This range for minimum pore
pressure increase required to cause fault reactivation is given for
critically oriented fault planes which are not observed at the Nay-
lor Field. Nevertheless, the pore pressure increase requiredto cause
fault reactivation for known fault planes at the Naylor Field ranges
from 2MPa to 17MPa (estimated pore pressure of 19.534.5MPa)
(Fig. 12).
The analytical model presented in this paper is useful to provide
an initial estimate of the stress changes with simplied reservoir
geometry and an assumed uniformpore pressure distribution. This
provides an easy way to perform sensitivity analysis. To better
infer the in situ stress changes with heterogeneous poromechan-
ical properties (various geological facies) with accurate reservoir
geometry and withthe modelled pore pressure change distribution
within the reservoir, it is essential to perform a 3D geomechanical
modelling.
In fact, if the pore pressure change within the reservoir is highly
localized, the induced stress changes may be signicant in the
bounding seal. The analysis detailed in this paper is focused on the
reservoir itself but it is also relevant to seal integrity. If faults are
reactivated or shear fractures are induced in the reservoir, they
could potentially propagate through the seal, thereby compromis-
838 S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839
Fig. 12. Fault reactivationpropensityfor all faults (a) usingscenario1: SSFRassump-
tion, (b) usingscenario2: NFRassumptionand(c) usingscenario3: SSFRassumption.
ing its hydraulic integrity. Due to the high permeability of Waarre C
Formation, andthe relative modest injectionrates (averaging about
124 tonnes per day), there are not large pressure gradients in the
near wellbore region, and so there is no risk of compromising seal
integrity near the injection well.
Fig. 13. Pore pressure prole simulated at Naylor fault.
There is also the concern that thermal stresses from injected
CO
2
colder than the formation might affect seal integrity near the
injection well. Temperature measurements fromdownhole gauges
in CRC-1 indicate that the injected gas at the reservoir level is about
20

C cooler than the reservoir temperature. However the comple-


tion interval in CRC-1 is at some distance below the main top seal,
the Belfast mudstone, and non-isothermal simulations of injection
indicate that the temperature change is only signicant very close
to the well. The possible effect of thermal stresses is still being
examined, but so far there is no evidence to suggest any signicant
changes in permeability near the well.
Between March 2008 and August 2009, 65,445 tonnes of CO
2
-
richgas were injected at CRC-1, witha pore pressure increase at the
CRC-1 well location of approximately 1.5MPa (Fig. 4). The reservoir
simulation updated with the latest acquired data (Fig. 13) esti-
mates a pore pressure of 18.5MPa at the Naylor Fault at the base of
the reservoir where the reactivation propensity is highest. In some
studied scenarios, the minimum pore pressure required to reac-
tivate a fault is 18.5MPa at the top of the reservoir and roughly
19MPa at the bottom of the reservoir. The microseismic array
deployed in the Naylor-1 well has recorded microseismic events
that may be attributable to fault movement, but the spatial uncer-
tainty exceeds 200m. Careful monitoring will help us to improve
our understanding of reservoir behaviour. In addition, more in situ
stress measurements will allowus to discriminate between the dif-
ferent assumptions that are made about the stress eld regime, the
fault strength and the reservoir stress path.
7. Conclusion
When assessing the suitability of possible CO
2
storage sites,
it is important to evaluate whether injection-related uid pres-
sure increases could reactivate pre-existing faults and generate
new fractures. Such brittle deformation could increase permeabil-
ity and promote unwanted movement of CO
2
out of the intended
storage area. Thus, in order to evaluate the fault reactivation
propensity during injection, a geomechanical analysis of the Nay-
lor Field, Otway Basin, Victoria was undertaken. Appropriate stress
orientations and gradients were determined from eld data, with
laboratory testing providing geomechanical properties of reservoir
rock samples. The stress regime in the onshore Victorian Otway
Basin was assumed to be strikeslip if maximum horizontal stress
is estimated using frictional limit and DITF occurrence but normal
if maximum horizontal stress is determined by sonic log inversion
fromCRC-1well. As aresult of theconictingdataandother geome-
chanical uncertainties, we performed sensitivity analyses using
S. Vidal-Gilbert et al. / International Journal of Greenhouse Gas Control 4 (2010) 827839 839
both stress elds and a range of potential geomechanical property
inputs.
Injection of CO
2
into the reservoir results in some perturba-
tion to the pore pressure prole and thus to some alterations in
the in situ stress eld acting on the reservoir and on its close
surroundings. Some analytical models were used to estimate the
stress paths followed by the in situ stresses. The geomechani-
cal model and the reservoir stress paths were used to estimate
the maximum sustainable pressure to avoid fault slip at injec-
tion site. The highest reactivation propensity (for critically oriented
faults) ranges from an estimated pore pressure increase (^P
p
) of
115.7MPa (estimated pore pressure of 18.533.2MPa) depending
on assumptions made about maximum horizontal stress mag-
nitude (SSFR with o
hmin
=14.5MPa/km and o
Hmax
=26MPa/km;
NFR with o
hmin
=14.5MPa/km and o
Hmax
=18MPa/km; SSFR with
o
hmin
=18.5MPa/km and o
Hmax
=37MPa/km), fault strength (C=0
and=0.6; C=5MPa and=0.76), reservoir stress path(=0, con-
stant horizontal stress and =0.4) and Biots coefcient (=0.7
and =1) and considering the magnitude of the minimum in situ
stress at the threshold for the injection pressure. For the fault
known at Naylor Field, the critical pore pressure changes range
from an estimated pore pressure increase of 217MPa (estimated
pore pressure of 19.534.5MPa).
The geomechanical model illustrates the important role of Biots
coefcient and reservoir stress path in controlling the risk of fault
reactivation. Availability of eld data and in situ stress measure-
ments obtained during the CO
2
injection within the reservoir will
improve and constrain the accuracy of the geomechanical model.
Acknowledgements
The authors acknowledge the funding provided to the Coopera-
tive ResearchCentre for Greenhouse Gas Technologies (CO2CRC) by
the Commonwealthof Australia toenable this researchtobe under-
taken. The authors also wish to acknowledge three other CO2CRC
contributors: Ric Daniel from the Australian School of Petroleum,
University of Adelaide, Australia for fruitful discussions on seal
capacity of caprocks, Tess Dance from CSIRO Petroleum Resources,
Perth, Australia for providing the geological model and Andy Nicol
from GNS Science, New-Zealand for reviewing this paper. The sug-
gestions for paper improvement from the journal reviewers are
greatly appreciated.
References
Addis, 1997. The stress-depletion response of reservoirs. In: SPE Annual Technical
Conference and Exhibition, SPE 38720, San Antonio, Texas, October 58.
Allen, T.I., Gibson, G., Cull, J.P., 2005. Stress-eld constraints from recent intraplate
seismicity in southeastern Australia. Australian Journal of Earth Sciences 52,
217229.
Berard, T., Sinha, B.K., van Ruth, P., Dance, T., John, Z., Tan, C., 2008. Stress estima-
tion at the Otway CO2 storage site, Australia. In: SPE Asia Pacic Oil and Gas
Conference and Exhibition, SPE 116422, Perth, Australia, October 2022.
Bouteca, M., 1994. Contributions of poroelasticity to reservoir engineering: labora-
tory experiments, application to core decompression and implication in HP-HT
reservoir depletion, Paper SPE/ISRM 28093.
Byerlee, J.D., 1978. Friction of rocks. Pure and Applied Geophysics 116, 615626.
Castagna, J.P., Batzle, M.L., Kan, T.K., 1993. Rock physics the link rock properties
and AVOresponse. In: Castagna, J.P., Backus, M. (Eds.), Offset-Dependent Reec-
tivity Theory and Practice of AVO Analysis. Investigations in Geophysics, No.
8. Society of Geophysicists, Tulsa, Oklahoma, pp. 135171.
Daniel, R.F., 2007. CarbonDioxideSeal CapacityStudy. CRC-1, CO2CRCOtwayProject,
Otway Basin, Victoria. CO2CRC Report No: RPT07-0629.
Dickinson, J.A., Wallace, M.W., Holdgate, G.R., Gallacher, S.J., Thomas, L., 2002. Origin
andtimingof theMiocene-PlioceneunconformityinSoutheast Australia. Journal
of Sedimentary Research 72, 288303.
Eshelby, J.D., 1957. The determination of the elastic eld of an ellipsoidal inclu-
sion and related problems. Proceeding of the Royal Society, London A241, 376
396.
Gardner, G.H.F., Gardner, L.W., Gregory, A.R., 1974. Formation velocity and density
the diagnostic basics for stratigraphic traps. Geophysics 39, 770780.
Grasso, J.R., 1992. Mechanics of seismic instabilities induced by the recovery of
hydrocarbons. PAGEOPH 139 (3/4), 507.
Hawkes, C.D., McLellan, P.J., Zimmer, U., Bachu, S., 2004. Geomechanical factors
affecting geological storage of CO2 in depleted oil and gas reservoirs: risks and
mechanisms. In: Gulf Rocks 2004, the 6th North America Rock Mechanics Sym-
posium (NARMS): Rock Mechanics Across Borders and Disciplines, Houston,
Texas, June 59.
Hawkes, C.D., Bachu, S., Haug, K., Thompson, A.W., 2005. Analysis of in-situ stress
regime in the Alberta Basin, Canada, for performance assessment of CO2 geolog-
ical Sequestration sites. In: Fourth Annual Conference on Carbon Capture and
Sequestration DOE/NETL, May 25.
Hettema, M.H.H., Schutjens, P.M.T.M., Verboom, B.J.M., Gussinklo, H.J., 1998.
Production-induced compaction of sandstone reservoirs: the strong inuence
of eld stress. Paper SPE 50630.
Hillis, R.R., Sandiford, M., Reynolds, S., Quigley, M.C., 2008. Present-day stresses,
seismicity and neogene-to-recent tectonics of Asutraliass passive margins:
intraplate deformation controlled by plate boundary forces. In: Johnson, H.,
Dore, A.G., Gatliff, R.W., Holdsworth, R., Lundin, E.R., Ritchie, J.D. (Eds.), The
Nature and Origin of Compression in Passive Margins. Geological Society, Lon-
don, pp. 201204, Special Publications 306.
Hillis, R.R., 2001. Coupled changes in pore pressure and stress in oil elds and sedi-
mentary basins. Petroleum Geosciences 7, 419425.
Hillis, R.R., Meyer, J.J., Reynolds, S.D., 1998. The Australian stress map. Exploration
Geophysics 29, 420427.
Jones, R.M., Boult, P.J., Hillis, R.R., Mildren, S.D., Kaldi, J., 2000. Integrated hydrocar-
bon seal evaluation in the Penola Trough, Otway Basin. APPEA Journal 40 (1),
194212.
Lyon, P.J., Boult, P.J., Watson, M., Hillis, R.R., 2005. A systematic fault seal evaluation
of the Ladbroke Grove and Pyrus traps of the Penola Trough, Otway Basin. APPEA
Journal 45 (1), 459474.
Mildren, S.D., Hillis, R.R., Kaldi, J., 2002. Calibrating predictions of fault seal reacti-
vation in the Timor Sea. APPEA Journal 42, 187202.
Moos, D., Zoback, M.D., 1990. Utilization of observations of well bore failure to
constrain the orientation and magnitude of crustal stresses: application to con-
tinental, Deep Sea Drilling Project, an Ocean Drilling ProgramBoreholes. Journal
of Geophysical Research 95, 93059325.
Nelson, E, Hillis, R.R., Sandiford, M., Reynolds, S., Mildren, S., 2006. Present-day state-
of-stress of Southern Australia. APPEA Journal 46, 283305.
Rhett, D.W., Risnes, R., 2002. Predicting critical borehole pressure and critical
reservoir pore pressure in pressure depleted and repressurized reservoirs. In:
SPE/ISRM Rock Mechanics Conference, SPE/ISRM 78150, Irving, Texas, October
2023.
Rogers, C., van Ruth, P.J., Hillis, R.R., 2008. Fault reactivation in the Port Camp-
bell Embayment with respect to carbon dioxide sequestration, Otway Basin,
Australia. In: Johnson, H., Dore, A.G., Gatliff, R.W., Holdsworth, R., Lundin, E.R.,
Ritchie, J.D. (Eds.), The Nature and Origin of Compression in Passive Margins.
Geological Society, London, pp. 201204, Special Publications 306.
Root, R.S., Gibson-Poole, C.M., Lang, S.C, Streit, J.E., Underschultz, J., Ennis-King, J.,
2004. Opportunities for geological storage of carbon dioxide in the offshore
Gippsland Basin, SE Australia: an example fromthe upper Latrobe Group PEASA
Eastern Australasian Basins Symposium II.
Rudnicki, J.W., 1999. Alteration of regional stress by reservoirs and other inhomo-
geneities: stabilizing or destabilizing? In: Proc. 9thIntern., Congress Rock Mech.,
vol. 3, Paris, France, 2528 August.
Sandiford, M., Wallace, M.W., Coblentz, D., 2004. Origin of the in situ stresss eld in
southeastern Australia. Basin Research 16, 325338.
Santarelli, F.J., Tronvoll, J.T., Svennekjaer, M., Skeie, H., Henriksen, H., Bratli, R.K.,
1998. Reservoir Stress-Path: The Depletion and the Rebound, paper SPE/ISRM
47350.
Schneider, C.L., Hill, K.C., Hoffman, N., 2004. Compressional growth of the Minerva
Anticline, Otway Basin, Southeast Australia evidence of oblique rifting. APPEA
Journal 44, 463480.
Segall, P., 1989. Earthquake triggered by uid extraction. Geology 17, 942946.
Sibson, R.H., 1974. Frictional constraints onthrust, wrenchandnormal faults. Nature
249, 542544.
Spencer, L., Xu, J., LaPedalina, F., Weir, G., 2006. Site Characterisation of the Otway
Basin Carbon Dioxide Geo-sequestration Pilot Project in Victoria, Australia. In:
8th International Conference on Greenhouse Gas Control Technologies, Trond-
heim, Norway, 1922 June (Abstract).
Streit, J.E., Hillis, R.R., 2004. Estimating fault stability and sustainable uid pressures
for underground storage of CO
2
in porous rock. Energy 29 (910), 14451456.
Van Ruth, P., 2007. Extended Leak Off test Report. CO2CRC Report No.: RPT07-0608.
Vidal-Gilbert, S., Nauroy, J.-F., Brosse, E., 2009. 3D geomechanical modelling for
CO
2
geologic storage in the Dogger carbonates of the Paris Basin. International
Journal of Greenhouse Gas Control 3 (3), 288299.
Wang, Z., 2000. Dynamic versus static elastic properties. In: Seismic and Acoustic
Velocities in Reservoir Rocks. SEG Geophysics Reprint Series, No. 19.
Watson, M., Boreham, C.J., Tingate, P., 2004. Carbon dioxide and carbonate cements
in the Otway Basin: implications for geological storage of carbon dioxide. APPEA
Journal 45, 703720.
Watson, M.N., Gibson-Poole, C.M., 2005. Reservoir selection for optimised geo-
logical injection and storage of carbon dioxide: a combined geochemical and
stratigraphic perspective. In: Conference Proceedings of the Fourth Annual Con-
ference on Carbon Capture and Sequestration DOE/NETL, Canada, May 25.
Zoback, M.D., 2008. Reservoir Geomechanics. Cambridge University Press.

Potrebbero piacerti anche