You are on page 1of 8

Tectonophysics 703–704 (2017) 1–8

Contents lists available at ScienceDirect

Tectonophysics

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

Lithology-dependent minimum horizontal stress and in-situ


stress estimate
Yushuai Zhang a, Jincai Zhang b,c,⁎
a
Wanfang College of Science and Technology, Henan Polytechnic University, Jiaozhuo, Henan 454000, China
b
College of Earth Science and Engineering, Shandong University of Science and Technology, Qingdao, Shandong, China
c
Geomech Energy, Houston, TX, USA

a r t i c l e i n f o a b s t r a c t

Article history: Based on the generalized Hooke's law with coupling stresses and pore pressure, the minimum horizontal stress is
Received 5 December 2016 solved with assumption that the vertical, minimum and maximum horizontal stresses are in equilibrium in the
Received in revised form 2 March 2017 subsurface formations. From this derivation, we find that the uniaxial strain method is the minimum value or
Accepted 4 March 2017
lower bound of the minimum stress. Using Anderson's faulting theory and this lower bound of the minimum hor-
Available online 6 March 2017
izontal stress, the coefficient of friction of the fault is derived. It shows that the coefficient of friction may have a
Keywords:
much smaller value than what it is commonly assumed (e.g., μf = 0.6–0.7) for in-situ stress estimate. Using the
Minimum horizontal stress derived coefficient of friction, an improved stress polygon is drawn, which can reduce the uncertainty of in-situ
In-situ stress stress calculation by narrowing the area of the conventional stress polygon. It also shows that the coefficient of
Coefficient of friction friction of the fault is dependent on lithology. For example, if the formation in the fault is composed of weak
Weak fault and stability shales, then the coefficient of friction of the fault may be small (as low as μf = 0.2). This implies that this fault
Improved stress polygon is weaker and more likely to have shear failures than the fault composed of sandstones. To avoid the weak
fault from shear sliding, it needs to have a higher minimum stress and a lower shear stress. That is, the critically
stressed weak fault maintains a higher minimum stress, which explains why a low shear stress appears in the
frictionally weak fault.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction indicate fracture closure pressures (i.e., the minimum stress). There-
fore, challenges still exist for efficiently, routinely and accurately de-
In-situ stress is one of the most important parameters for geosci- termining the minimum horizontal stress. Similarly, challenges also
ence modeling and geo-engineering design, particularly in the oil exist in theoretically estimating the horizontal stresses (e.g.,
and gas industry. For example, the minimum horizontal stress is Schmitt et al., 2012; Zhang, 2013a).
very critical for fracture gradient prediction, casing design and In order to constrain in-situ stresses, the stress polygon based on
wellbore stability assessment in drilling operations and for planning stress regimes has been used for decades (e.g., Zoback et al., 2003;
hydraulic fracturing in tight reservoirs. The in-situ stress includes Sibson, 1974); however, it needs input of the coefficient of friction
three mutually orthogonal principal stresses in the subsurface; typi- of the fault which may not be easily obtained. Conventionally, it is as-
cally, they are the vertical (overburden) stress, the maximum hori- sumed that the coefficient of friction of the fault is a constant (μf =
zontal and minimum horizontal stresses (σV, σH, σh). In-situ stress 0.6–0.7) based on Byerlee's law (Byerlee, 1978). This assumption
can be measured from borehole coring samples using the anelastic may cause a large uncertainty in in-situ stress estimation because
strain recovery method (e.g., Cui et al., 2014). The minimum hori- of the uncertainty in estimate of μ f ; in fact, some faults are weak
zontal stress is commonly estimated from downhole tests, such as with very low μ f . For instance, Bird and Kong (1994) and Carena
leak-off test (LOT), extended leak-off test (XLOT), or diagnostic frac- and Moder (2009) concluded that all faults in the vicinity of the
ture injection test (DFIT) (e.g., Bredehoeft et al., 1976; Haimson and transform plate boundary of the western USA are frictionally weak
Fairhurst, 1967; Raaen et al., 2006; Zhang et al., 2008; Lang et al., to very weak (μf ≤ 0.2). Iaffaldano (2012) inferred that the coefficient
2011). However, LOT data normally are not available at the location of friction of large-scale plate boundaries is in the range of 0.01–0.07.
and depth of interest. Furthermore, most LOT tests only provide frac- Byerlee (1978) pointed out that if the sliding surfaces are separated
ture initiation pressures or fracture breakdown pressures and do not by gouge composed of some clay minerals, the friction is very low.
Engelder and Fischer (1994) concluded that the minimum horizontal
⁎ Corresponding author at: Geomech Energy, Houston, TX, USA. stress calculated from a friction of μf = 0.6 underestimates the min-
E-mail address: jjczhang@163.com (J. Zhang). imum stress in the central North Sea Graben and does not match the

http://dx.doi.org/10.1016/j.tecto.2017.03.002
0040-1951/© 2017 Elsevier B.V. All rights reserved.
2 Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8

leak-off data in the Scotian Shelf, Canada. Not surprisingly, recent using the principal effective stresses is shown in the following equation
work shows that the coefficient of friction of the fault is highly relat- when there is no shear failure/sliding in the fault:
ed to the lithology or mineralogy of the fault gouge. For example, ex-
treme fault weakness (μf ~ 0.1) occurs within a 3 m wide creeping 2C cosϕ f 1
fault core (Zoback et al., 2010) in the San Andreas of central Califor- σ 01 ≤ þ σ0 ð1Þ
1− sinϕ f k f 3
nia because of the presence of weak clay minerals (Carpenter et al.,
2011; Collettini et al., 2011). Saffer and Marone (2003) observed a
1− sinϕ f qffiffiffiffiffiffiffiffiffiffiffiffiffiffi −2
coefficient of friction in the fault gouge of 0.42–0.68 for illitic shale;
kf ¼ ¼ μ 2f þ 1 þ μ f ð2Þ
however, under identical conditions, a low friction (μ f = 0.15– 1 þ sinϕ f
0.32) is inferred in a smectitic shale. There are also questions
concerning whether μf is the same for all faults in a region, whether
where σ′1 is the maximum effective stress, σ′3 is the minimum effective
it is even constant along strike on the same fault (Carena and
stress, C is the cohesive strength of the fault, φf is the angle of internal
Moder, 2009), or whether it is depth-dependent. Field measure-
friction of the fault, μf is the coefficient of friction of the fault, and μf =
ments of in-situ stress shows that the horizontal stresses are highly
tanϕf.
dependent on lithologies (e.g., Warpinski and Teufel, 1989;
For deep formations, the cohesion of the fault is often neglected.
Wileveau et al., 2007; Gunzburger and Magnenet, 2014). For in-
Therefore, to avoid a fault from frictional sliding, the in-situ stress
stances, coal seams and shales have much higher minimum horizon-
should satisfy the following equation - a similar equation was used by
tal stresses than the adjacent sandstones, as shown in Fig. 1.
Zoback et al. (2003) and Sibson (1974):
The commonly-used methods for determining in-situ stress are
briefly reviewed in this paper. Then, an improved method for calculat-
1 0
ing the minimum horizontal stress is given. Because an in-situ stress σ 01 ≤ σ ð3Þ
polygon is closely related to the coefficient of friction of the fault, this kf 3
parameter needs to be accurately determined. A new method is given
to derive the coefficient of friction of the fault. This method can explain In porous media, the effective stress and total stress have the follow-
why the horizontal stresses depend on lithologies. ing relationship (Biot, 1941):

1.1. In-situ stress in various faulting regimes


σ 0ij ¼ σ ij −αδij pp ð4Þ
Three stress regimes from Anderson's faulting theory (Anderson,
1951) can be used to describe in-situ stress states (e.g., Zoback et al., where σ′ij and σij are the effective stress and total stress tensors, respec-
2003; Peng and Zhang, 2007): tively, α is Biot's effective stress coefficient, δij is the Kronecker delta, pp
is the pore pressure.
1. Normal faulting stress regime. In this case, the vertical stress drives It should be noted that pore pressure is not hydrostatic in many deep
normal faulting, and fault slip occurs when the minimum stress formations, but overpressured, as shown in Fig. 2. To accurately deter-
reaches a sufficiently low value. In this stress state, the vertical stress mine in-situ stress, the first requirement is to obtain pore pressure of
is the greatest principal stress, i.e. σV ≥ σH ≥ σh. the formations (e.g., Eaton, 1975; Zhang, 2011). From Eqs. (3) and (4),
2. Strike-slip faulting stress regime. In this case, the vertical stress is the the in-situ stresses can be expressed in the following forms for different
intermediate principal stress, i.e. σH ≥ σV ≥ σh. faulting stress regimes:
3. Reverse (or thrust) faulting stress regime. In this case, the vertical
stress is the least principal stress, i.e. σH ≥ σh ≥ σV.
Assuming there are critically oriented faults constraining stress mag-
nitudes, the Mohr-Coulomb criterion (Jaeger and Cook, 1979) for a fault

Fig. 2. In-situ stress and pore pressure (pp and MDT) versus depths in a deepwater well
with the lower bound of the minimum horizontal stress (Sh_LB) and upper bound of
the maximum horizontal stress (SH_UB) calculated from Eqs. (8) and (9) assuming a
Fig. 1. The minimum horizontal stress profile (σh) measured from min-frac tests in constant μf of 0.6. The MDT points are the measured formation pore pressures from the
Colorado (modified from Warpinski and Teufel, 1989). Note 1 ft. = 0.305 m, 1 psi = borehole after drilling; Sv is the vertical stress; LOT data are the measured formation
0.00689 MPa. leak-off pressures.
Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8 3

Normal faulting regime: Fig. 2, where the vertical stress and measured pore pressure at depth of
4316 m from the sea level (3458 m from the sea floor) are: σV =
σ 01 σ V −αpp 1 82.5 MPa, pp = 69.7 MPa. For simplification, a Biot's coefficient of 1 is ap-
¼ ≤ ð5Þ
σ 03 σ h −αpp k f plied to Eqs. (8) and (9) to calculate the bounds of the minimum and
maximum horizontal stresses. From these data, an in-situ stress polygon
Strike-slip faulting regime: is plotted, as shown in Fig. 3. The maximum and minimum horizontal
stresses are constrained in the stress polygon by considering the three dif-
σ 01 σ H −αpp 1 ferent stress regimes. However, there is still a large uncertainty. Therefore,
¼ ≤ ð6Þ
σ 03 σ h −αpp k f the stress polygon may need improvement.

Reverse faulting regime: 1.2. Vertical stress

σ 01 σ H −αpp 1
¼ ≤ ð7Þ Vertical stress is generated by the weight of the overlying forma-
σ 03 σ V −αpp k f tions. Hence, it can be obtained by integrating bulk density logs, which
are normally available in exploration and production wells. Therefore,
Hence, from Eq. (5) the lower bound of the minimum horizontal vertical stress can be calculated by the following equation:
stress (σLB
h ) can be obtained:
Z z
 
σ V ¼ ρw gzw þ g ρb ðzÞdz ð10Þ
σ LB
h ¼ k f σ V −αpp þ αpp ð8Þ zw

From Eq. (7), the upper bound of the maximum horizontal stress where ρb (z) is the formation bulk density as a function of depth and can
(σUB
H ) can be expressed as: be obtained from well logging (Zhang, 2013b), ρw is the density of sea
water for offshore drilling, z is the depth from the sea level, zw is the
1  water depth, and in onshore case, zw = 0.
σ UB
H ¼ σ V −αpp þ αpp ð9Þ
kf
1.3. The maximum horizontal stress
It should be noted that Biot's effective stress coefficient, α, ap-
proaches to 1 when the rocks approach to failure states. When a borehole is drilled with a mud pressure higher than the for-
Fig. 2 plots in-situ stress profile estimated from the above method in mation breakdown pressure, the drilling-induced tensile fracturing oc-
a deepwater well with water depth of 858 m. The figure shows the pore curs. The maximum horizontal stress in the case of non-penetrating
pressure, vertical stress and the calculated lower and upper bounds of fluid (for impermeable formations) can be estimated from drilling
the horizontal stresses (Eqs. (8) and (9)) to constrain in-situ stress. In induced tensile fractures in a vertical well using the following elastic
the normal faulting stress regime, the minimum horizontal stress solution (originally derived by Haimson and Fairhurst, 1967), if a dril-
should be in between of the lower bound of the minimum horizontal ling-induced tensile fracture occurs at a mud pressure of pw:
stress and the overburden stress. However, in the strike-slip and thrust
stress regimes, the maximum horizontal stress should be located in be- σ H ¼ 3σ h −pw −pp þ T 0 þ σ t ð11Þ
tween of the overburden stress and the upper bound of the maximum
horizontal stress. It can be observed that the range of the lower bound where pw is the downhole mud pressure when tensile fracturing is in-
of minimum horizontal stress and the upper bound of the maximum duced, T0 is the tensile strength of the rock, σt is the thermal stress, σ t ¼
horizontal stress is so big. Therefore, the uncertainty of the in-situ stress αt E
1−ν ðT w −T f Þ; Tw, Tf are the downhole mud and formation temperatures,
estimation is very large. respectively.
The stress polygon can be drawn using the in-situ stress and pore Detournay and Cheng (1988) proposed a poroelastic solution for
pressure relationships in different stress regimes from Eqs. (5)–(9). This breakdown pressure induced by pressurization of a borehole in hydrau-
is shown in Fig. 3 assuming μf = 0.6. Fig. 3 uses the example shown in lic fracturing in the case of penetrating fluid (for permeable formations).
The maximum horizontal stress in permeable formations can be derived
from it, if a drilling-induced tensile fracture occurs at a mud pressure of
pw in a vertical well:

σ H ¼ 3σ h −2ð1−ηÞpw −2ηpp þ T 0 þ σ t ð11aÞ

where η is a poroelastic coefficient ranging from 0 to 0.5 and


η = α(1− 2ν)/[2(1− ν)], ν is the drained Poisson's ratio.
The maximum horizontal stress can also be estimated from wellbore
breakouts when the applied mud weight in drilling operations is less
than the shear failure gradient (Barton et al., 1988). Li and Purdy
(2010) proposed a new method to determine the maximum horizontal
stress using observations of the angle of the width of the breakout in
vertical boreholes:

UCS þ ðq þ 1Þpw −α ðq−1Þpp −ð1−2 cos2βb Þσ h þ σ t


σH ≤ ð12Þ
1 þ 2 cos2βb

Fig. 3. Stress polygon obtained from Eqs. (8) and (9) with measured pore pressure of
where UCS is the uniaxial compressive strength of the rock, 2ßb is the
69.7 MPa and vertical stress of 82.5 MPa in a borehole, assuming that the coefficient of
friction of the fault (μf) is 0.6. The plotted depth is 4316 m TVDss or 3458 m below the wellbore breakout angle, pw is the mud pressure, q is a parameter
sea floor (refer to Fig. 2). In the plot, NF, SS, and RF represent the normal, strike-slip, and related to the internal friction angle of the rock with q = (1 + sin φ)/
reverse faulting stress regimes, respectively. (1− sin φ), ϕ is the angle of internal friction of the rock.
4 Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8

2. Commonly-used methods for determining the minimum horizon- curve remains fairly linear until formation breakdown, as shown in
tal stress Fig. 4(b). In this case the fracture initiation pressure (pi) is the break-
down pressure (pb). It should be noted that a term of leak-off pressure
2.1. Minimum stress measurements from the leak-off tests (LOP) is normally used as the LOT pressure. The leak-off pressure is de-
fined as that a pressure at a point in a pressure-time plot deviates from
The minimum horizontal stress can be determined by direct mea- linearity (Edwards et al., 1998). Therefore, the leak-off pressure or LOT
surements, i.e. via the commonly accepted method of micro-hydraulic pressure can have a value anywhere between the fracture initiation
fracturing (e.g., Haimson and Cornet, 2003), or its oil field equivalent: pressure (pi, as shown in Fig. 4(a)) and the breakdown pressure (pb,
mini-frac, leak-off test (LOT), or extended leak-off test (XLOT, White as shown in Fig. 4(b)). It is clear that the leak-off pressure is not the min-
et al., 2002). The minimum stress in normal faulting stress regime is imum horizontal stress. Some empirical correlations between the leak-
the minimum principal horizontal stress. This stress is typically equal off pressure (pLOP) and the minimum horizontal stress are obtained
to the fracture closure pressure, which can be observed after shut-in from leak-off tests, such as σh = 0.901pLOP (Breckels and van Eekelen,
on the decline curve of a leak-off test (Whitehead et al., 1986; Zhang 1982; Edwards et al., 1998).
and Roegiers, 2010). Fig. 4 schematically presents two typical cases of The minimum horizontal stress and tensile strength can be obtained
extended leak-off tests, each case with two pressurization cycles. when XLOT data are available; then, the maximum horizontal stress in
In a typical leak-off test, a borehole interval is isolated and sealed both impermeable and permeable cases can be calculated from the
with the casing or inflatable packers; i.e., only the open hole below breakdown pressure using Eqs. (11) and (11a) by replacing pw with pb.
the casing and any new formation drilled prior to the test are exposed
to fluid pumped at a constant rate. The pressure increase in the hole is 2.2. Minimum stress estimation from the uniaxial strain model
typically linear as long as there are no leaks in the system, and the ex-
posed formation is not highly permeable. At some point, the rate of Another commonly-used method to estimate the minimum horizon-
pressurization changes such that the pressure-time curve departs tal stress is based on a uniaxial strain assumption. This method assumes
from linearity. This departure from linearity is referred to as the fracture that the two horizontal stresses are equal and that no lateral strain exists.
initiation pressure (pi), as shown in Fig. 4(a). The pressure is then typi- The minimum horizontal stress obtained from the uniaxial strain model
cally seen to increase at a lower rate until a maximum pressure, the can be expressed as a function of vertical stress and pore pressure:
breakdown pressure (pb), is reached. After the rock is broken down (hy-
draulic fracture created), at some point the pressure in the borehole re- ν  
σh ¼ σ V −αpp þ αpp ð15Þ
mains fairly constant (pprop) at the same flowrate; the fracture is min
1−ν
propagating. When the pumps are turned off, the pressure immediately
drops to the instantaneous shut-in pressure (pisip). After the well is where σh_min is the minimum horizontal in-situ stress, ν is the Poisson's
shut-in, the pressure begins to decline as the fracture starts to close; ratio and can be obtained from the compressional and shear velocities
the pressure acting to close the fracture is equal to the minimum hori- ðvp =vs Þ2 −1
1

zontal stress (σh); i.e.: (vp and vs), and ν ¼ 2 .


ðvp =vs Þ2 −1
However, it has been found that this method or Eq. (15) is not accu-
σ h ¼ pc ð13Þ rate for estimating the minimum stress (e.g., Whitehead et al., 1986;
Whitehead et al., 1987; Meng et al., 2011). In fact, this equation is only
where pc represents the closure pressure. a special case of the minimum horizontal stress, because horizontal
A second pressurization cycle can be performed to obtain more data. strains normally are not zeros, and σh ≠ σH. This is one of the reasons
Because a fracture has been created by the first cycle, there is no tensile why it is over-simplified.
strength in the fracture reopening; i.e.:
3. The minimum horizontal stress calculation from in-situ stress
T 0 ¼ pb −pr ð14Þ configuration

where pb is the fracture breakdown pressure, pr is the fracture 3.1. Calculation of the minimum horizontal stress
reopening pressure.
In some case, particularly in brittle rocks or rocks with high tensile It could be assumed that the three principal stresses in subsurface
strength, no fracture initiation pressure is seen, and the pressure-time formations might satisfy Hooke's law. According to Hooke's Law, the

Fig. 4. Schematic plots of the pumping pressure and time or volume of injected fluid in two extended leak-off tests. (a) Typical XLOT with fracture initiation pressure; (b) XLOT without
fracture initiation pressure (or fracture initiation pressure is equal to formation breakdown pressure) and with high tensile strength.
Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8 5

minimum horizontal strain can be written as the following form, if the minimum horizontal stress, pore pressure, and Poisson's ratio are avail-
stresses are expressed in effective stress forms: able. Using Eq. (20), the minimum value of the minimum horizontal
  stress at different depths is calculated by assuming α = 1. The calculated
σ 0h −ν σ 0V þ σ 0H values and measured data are plotted in Fig. 5. The calculated minimum
εh ¼ ð16Þ
E values of the minimum horizontal stresses obtained from the proposed
method (Eq. (20)) are compared to the measured results and are
where εhis the strain in the minimum horizontal stress direction, E is the
shown in Fig. 5. It indicates that the uniaxial strain model or Eq. (20) is
Young's modulus, σV′, σH′ and σh′ are the effective vertical, maximum hor-
the lower limit of the minimum horizontal stress. The measured data
izontal and minimum horizontal stresses, respectively, ν is the Poisson's
also show that the sandstone has a lower minimum horizontal stress
ratio.
than the shale. This may be caused by the fact that sandstones normally
Solving Eq. (16) one obtains:
have lower values of Poisson's ratios than those in shales.
σ 0h −νσ 0H ¼ Eεh þ νσ 0V ð17Þ
4. Lithology-dependent coefficient of friction and improved method
Because σH′ ≥ σh′, σh′ − νσH′ ≤ σh′ − νσh′. Hence, the following relation- for in-situ stress estimate
ship can be obtained from Eq. (17):
4.1. Lithology-dependent coefficient of friction of the fault
ν E
σ 0h ≥ σ0 þ ε ð18Þ
1−ν V 1−ν h Studies and experiments (e.g., Takahashi et al., 2007; Tembe et al.,
2010) in clay-quartz gouges show that the clay content has a significant
Normally, the layered formations extend very long in two horizontal
effect on the frictional strength of the fault; i.e., as clay content increases,
directions; therefore, the strain in the minimum horizontal direction is
the coefficient of friction decreases (e.g., when clay content is 100%,
much smaller than the strains in the vertical direction. Particularly,
μf b 0.1 for smectite). Analyzing the data of Takahashi et al. (2007), the
when the formations of interest are laterally constrained by stiff faults,
following linear relationship can be obtained for a smectite and quartz
the stress state is similar to the condition of uniaxial strain loading. In
mixture:
this extreme case, εh is close to zero. Therefore, from Eq. (18) the mini-
mum value of the effective minimum horizontal stress can be expressed
as the following: μ f ¼ 0:68−0:6C S ð23Þ
ν
σ 0h ¼ σ0 ð19Þ
1−ν V
where CS is the weight fraction of clay content (smectite) and CS is be-
Substituting Eq. (4) into Eq. (19), the minimum value of the mini- tween 0 and 1.
mum horizontal stress can be written as following: Analyzing the data from Tembe et al. (2010), the following linear re-
lationship can be obtained for illite-quartz mixture:
ν  
σ min
h ¼ σ V −αpp þ αpp ð20Þ
1−ν
μ f ¼ 0:68−0:42C I ð24Þ
This is the same form as the one obtained from the uniaxial strain
model (i.e., Eq. (15)). However, it is not a general case for calculating
the minimum horizontal stress, but only a special case - the minimum where CI is the weight fraction of clay content (illite) and CI = 0–1.
value of the minimum horizontal stress. Therefore, the coefficient of friction of the fault is lithology-depen-
Combining Eqs. (4) and (18) the minimum horizontal stress can be dent, and this needs to be considered when one applies Eqs. (8) and
written in the following form: (9) to estimate in-situ stress.

ν   E
σh ≥ σ V −αpp þ αpp þ ε ð21Þ
1−ν 1−ν h

From Eq. (21), the minimum horizontal stress can be estimated


when the vertical stress, pore pressure, Poisson's ratio and horizontal
strain are known. It can be seen that the uniaxial strain model (Eq.
(15)) is only a specific case of the minimum horizontal stress.
Since the horizontal strain (εh) is difficult to determine, the follow-
ing simplified equation can be used to estimate the minimum horizontal
stress, i.e.:

ν   c
σh ¼ σ V −αpp þ αpp þ σV ð22Þ
1−ν 1−ν

where c is the minimum stress coefficient and can be obtained by cali-


brating in-situ measured data (e.g., LOT, XLOT, DFIT).

3.2. Application of the proposed minimum horizontal stress method

The proposed method is applied to calculate the minimum horizontal


stress in the Piceance basin near Rifle of Colorado, USA. Extensive in-situ
stress measurements were performed in this basin by small-volume, Fig. 5. Vertical stress, pore pressure and minimum stress measured from the hydraulic
hydraulic-fracture stress tests (Warpinski et al., 1985). We examine the fracture tests in Colorado (Warpinski et al., 1985). The calculated minimum values of
test results and well logging data in this field. The vertical stress, the minimum horizontal stress (Sh-min) from Eq. (20) are plotted for comparison.
6 Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8

4.2. Relationship of the coefficient of friction of the fault and Poisson's ratio estimated that the slip on fault surfaces is consistent with μf = 0.11 ±
0.02.
As indicated, two methods can be used to calculate the lower bound Therefore, the common assumption of a fault strength of μf = 0.6–
of the minimum horizontal stress: one from a uniaxial strain model (Eq. 0.7 may be inappropriate. From Eq. (27), the case of μf = 0.6–0.7
(20)), the other from a stress regime constraint (Eq. (8)). Proximity that would corresponds to a low Poisson's ratio rock (e.g., sandstone, lime-
the two lower bounds are equal, the coefficient of friction can be esti- stone). For shales and other ductile rocks, μf values should be lower be-
mated, as shown in the following: cause of the higher values of Poisson's ratios (e.g., if ν = 0.4, μf = 0.2).
Therefore, the fault strength is lower when the formations mainly con-
ν sist of shales. Thus, the strength or the coefficient of friction of the fault
kf ¼ ð25Þ
1−ν may not be as large as has been previously assumed (i.e., μf = 0.6–0.7). It
should be smaller (e.g., as small as μf = 0.1–0.2) in shales or mudrocks.
Substituting Eq. (2) to Eq. (25), one obtains:
Therefore, using the same μf for different rocks may be inappropriate be-
sinϕ f ¼ 1−2ν ð26Þ cause it may not reflect the real properties of the rocks and faults.

4.3. The minimum horizontal stress calibration and stress polygon


1−2ν
μ f ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð27Þ constraint
2 ν ð1−νÞ
Using the coefficient of friction calculated from Eq. (27) or Eqs.
This equation is applicable in the following range of Poisson's ratio: (28)–(29), a stress polygon can be prepared, and the uncertainty of
0.16 b ν b 0.5. By submitting Eq. (25) to Eqs. (8) and (9) the lower the minimum horizontal stress can be reduced. Using the same case
bound of the minimum horizontal stress and the upper bound of the study as shown in Fig. 2, the in-situ stress can be better constrained by
maximum horizontal stress can be obtained as following. the aforementioned procedure with the coefficient of friction (μf)
In normal and strike-slip faulting regimes: being calculated from Poisson's ratio using Eq. (27) or directly using
ν   the new lower bound and upper bound horizontal stresses from Eqs.
σ LB
h ¼ σ V −αpp þ αpp ð28Þ (28) and (29). That is shown in Fig. 6 and it presents the lower and
1−ν
upper bounds of the horizontal stresses calculated for μf derived from
In strike-slip and reverse faulting regimes or tectonic stress regime: Eq. (27) at the depths from 3200 to 6800 m. The formations are mainly
shales from 3200 to 5800 m with μf of 0.15–0.3 (if Eq. (27) is used). This
1−ν   corresponds to a higher minimum horizontal stress. At the depth of
σ UB
H ¼ σ V −αpp þ αpp ð29Þ
ν N6000 m the formations are mostly sandstones. They have larger coef-
ficients of friction (0.4–0.56); therefore, a lower minimum horizontal
Hence, the proposed equation to estimate the minimum horizontal stress is calculated. Compared to Fig. 2, the range of the stress bounds
stress (Eq. (22)) can be rewritten as the following form: from the calculated coefficients of friction (Eq. (27)) in Fig. 6 is much
    smaller. Therefore, in-situ stress is better constrained and the uncertain-
σ h ¼ k f σ V −αpp þ αpp þ c 1 þ k f σ V ð30Þ ty in the stress estimation is reduced. Fig. 6 also implies that the coeffi-
cient of friction of the fault depends on Poisson's ratio or lithology,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi −2 because of Poisson's ratio is a function of the lithology.
ν
where k f ¼ 1−ν or k f ¼ ð μ 2f þ 1 þ μ f Þ ; c can be obtained from ana- Fig. 7 shows the stress polygon for the same case (Fig. 3) with a co-
lyzing measured data; e.g., c = 0.024 and k0 = 0.45 in the Qinshui efficient of friction calculated from the proposed method or Eq. (27) (μf
coalbed basin (Meng et al., 2011). of 0.2). Compared the new stress polygon (μf of 0.2) to the old one (μf of
From Eqs. (28) and (29), the new stress polygon can be plotted, and 0.6), the new polygon is smaller and the two horizontal stresses are
this stress polygon depends highly on Poisson's ratio. Eq. (27) indicates much better constrained (Fig. 7). With this proposed method and com-
that the coefficient of friction of the fault can be related to Poisson's bining other methods (borehole breakouts and drilling-induced tensile
ratio. Poisson's ratio is dependent on lithology and depth; therefore, fractures; e.g., Bradley, 1979; Zhang et al., 2003a; Zhang and Roegiers,
the coefficient of friction of the fault should also depend on lithology 2005; Zhang et al., 2009), in-situ stress can be more accurately
and depth. For example, a sandstone normally has a smaller Poisson's estimated.
ratio than a shale; hence, the sandstone in the fault has a larger coeffi- Fig. 7 can be explicitly illustrated by Mohr circles, as shown in Fig. 8.
cient of friction (e.g., when ν = 0.23, μf = 0.64 from Eq. (27)) than The Mohr-Coulomb failure envelope and maximum shear stress (τfmax)
the shale. This may be the reason why the sandstone normally has a along the fault are:
smaller horizontal stress. A shale normally has a higher Poisson's ratio,
thus a smaller coefficient of friction (e.g., ν = 0.4, therefore μf = 0.2 τ f ¼ μ f σ 0n ð31Þ
from Eq. (27)). This is verified by experimental results; for example,
the measurements by Ikari et al. (2011) show that fault gouges contain- σ 0V −σ 0h
ing clay minerals are frictionally weak (μf b 0.5), whereas gouges rich in τf max ¼ ð32Þ
2
silicate minerals (e.g., quartz, feldspar) are stronger (μf N 0.6). Goulty
(2008) pointed out that the residual shear strength for any lithology is where σ′n is the effective normal stress, σ′ V and σ′ h are the maxi-
governed by the coefficient of friction on a fault surface after initial dis- mum and minimum effective stresses, respectively, τ f and τ fmax
placement has occurred, which is lower than the coefficient of friction are the shear stress and the maximum shear stress in the fault,
measured by a peak strength test in a triaxial cell. The reduced coeffi- respectively.
cient of friction is known as the ‘coefficient of residual friction’ or the co- Fig. 8 indicates that in the critically stressed condition (before shear
efficient of friction of the fault. From measurements using a ring shear failure) a stronger fault (μf = 0.6 in Fig. 8) has a smaller effective mini-
apparatus, Bishop et al. (1971) found coefficient of residual friction is mum stress (σ′h_s), or higher shear stress (τs_max). However, a friction-
about 0.16 in undisturbed samples of London Clay, which is ally weak fault (e.g., μf = 0.2 in Fig. 8) needs a higher effective
stratigraphically equivalent to the Eocene claystones in the North Sea minimum stress (σ'h_w) (thus smaller Mohr circle and much lower
Basin. From well data in claystone sequences containing polygonal shear stress, τw_max) to keep the fault stability. This may potentially ex-
fault systems in the Central North Sea, Goulty and Swarbrick (2005) plain the low inferred shear stresses along strike-slip faults, such as the
Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8 7

Fig. 6. In-situ stress profile versus depth below the sea level for the same case as shown in Fig. 2. The in-situ stress, particularly the maximum horizontal stress, is better constrained
compared to Fig. 2. The left track plots Poisson's ratios derived from sonic logs; the middle track shows the coefficient of frictions calculated from the Poisson's ratio from Eq. (27); the
right track presents vertical stress, pore pressure, the minimum and maximum horizontal stress bounds calculated by the proposed method from Eqs. (28) and (29).

San Andreas (e.g., Hickman, 1991; Townend and Zoback, 2004). The link 5. Conclusions
between frictional strength and minimum stress has important implica-
tions for slip behavior on natural faults. For instance, in order to keep the A lower bound minimum horizontal stress can be estimated from
frictionally weak fault from slip, an unstable natural fault would require uniaxial strain model and Anderson's faulting theory. Combining these
modification of the minimum stress (increase of the minimum stress, as two methods, the coefficient of friction of the fault can be obtained
suggested in this paper) or the fault gauge composition (as suggested by using the formation Poisson's ratio. The derived coefficient of friction
Ikari et al., 2011). of the fault varies with depth and lithology. For a rock with a high
Poisson's ratio (such as some shales), the coefficient of friction of a
fault is calculated to be low; e.g., if ν = 0.4, then μf = 0.2. Therefore,
4.4. Lithology-dependent minimum horizontal stress the assumption of fault strength of μf = 0.6–0.7 may be only appropriate
for the rock (e.g., sandstone or limestone) with a low Poisson's ratio. For
To illustrate the relationship between lithology and the minimum the shales and other ductile rocks, the μf values should be lower because
horizontal stress, the measured minimum horizontal stress in Fig. 5 is of the higher value of Poisson's ratio (e.g., ν = 0.4, then μf = 0.2). There-
schematically plotted in Fig. 9. It can be observed that the sandstone fore, the fault strength is lower or the minimum horizontal stress is larg-
normally has a smaller minimum horizontal stress, because of its small- er when the formations mainly consist of shales (normally μf = 0.2). For
er Poisson's ratio. By contrast, the shale has a higher minimum horizon- a frictionally weak fault it needs a higher minimum stress (or much
tal stress; therefore, it can be used as a barrier of the hydraulic fracture lower maximum shear stress) to avoid the fault from slip. This may po-
propagation when hydraulic fracturing is performed in adjacent tentially explain the low inferred shear stresses along strike-slip faults,
sandstone. such as the San Andreas. Correlating the frictional coefficient with
Poisson's ratio has its engineering convenience, but may have uncer-
tainties. For the engineering applications, we recommend calibrating
the proposed method with measured data.

Fig. 8. Mohr circle diagrams showing the interaction of fault coefficient of friction, shear
Fig. 7. Stress polygons with the same vertical stress and pore pressure (as shown in Fig. 3) stress and effective stresses (maximum or vertical stress and minimum stress) in
in the shale at the depth of 4316 m below the sea level, but with two coefficients of critically stressed state. The vertical stress and pore pressure are the same to the one
frictions of the fault μf = 0.6 and μf = 0.2. The latter is calculated from the proposed shown in Figs. 3 and 7 with two coefficients of frictions of μf = 0.6 for a stronger fault
method (Eq. (27)), which has a much smaller stress polygon, i.e., a much lower and μf = 0.2 for a weak fault. The frictionally weak fault requires a much larger
uncertainty in stresses. minimum stress (or lower maximum shear stress) to maintain it stable.
8 Y. Zhang, J. Zhang / Tectonophysics 703–704 (2017) 1–8

Goulty, N.R., 2008. Geomechanics of polygonal fault systems: a review. Pet. Geosci. 14,
389–397.
Goulty, N.R., Swarbrick, R.E., 2005. Development of polygonal fault systems: a test of hy-
potheses. J. Geol. Soc. Lond. 162, 587–590.
Gunzburger, Y., Magnenet, V., 2014. Stress inversion and basement-cover stress transmis-
sion across weak layers in the Paris basin, France. Tectonophysics 617, 44–57.
Haimson, B.C., Cornet, F.H., 2003. ISRM suggested methods for rock stress estimation –
part 3: hydraulic fracturing (HF) and/or hydraulic testing of pre-existing fractures
(HTPF). Int. J. Rock Mech. Min. Sci. 40, 1011–1020.
Haimson, B.C., Fairhurst, C., 1967. Initiation and extension of hydraulic fractures in rocks.
SPE J. 7, 310–318.
Hickman, S.H., 1991. Stress in the lithosphere and the strength of active faults. Rev.
Geophys. 759–775 (U.S. Natl. Rep. to Int. Union of Geodesy and Geophys).
Iaffaldano, G., 2012. The strength of large-scale plate boundaries: constraints from the dy-
namics of the Philippine Sea plate since ~ 5 Ma. Earth Planet. Sci. Lett. 357–358,
21–30.
Ikari, M.J., Marone, C., Saffer, D.M., 2011. On the relation between fault strength and fric-
tional stability. Geology 1, 83–86.
Jaeger, J.C., Cook, N.G.W., 1979. Fundamentals of Rock Mechanics. third ed. John Wiley &
Sons.
Lang, J., Li, S., Zhang, J., 2011. Wellbore stability modeling and real-time surveillance for
deepwater drilling to weak bedding planes and depleted reservoirs. Paper SPE/
IADC 139708.
Fig. 9. Schematic representation of the relationship of lithology and the minimum Li, S., Purdy, C., 2010. Maximum horizontal stress and wellbore stability while drilling:
horizontal stress. modeling and case study. Paper SPE 139280.
Meng, Z., Zhang, J., Wang, R., 2011. In-situ stress, pore pressure and stress-dependent per-
meability in the Southern Qinshui Basin. Int. J. Rock Mech. Min. Sci. 48, 122–131.
We find and verify that the uniaxial strain method is the minimum Peng, S., Zhang, J., 2007. Engineering geology for underground rocks. Spring 2007.
value or lower bound of the minimum stress. This minimum stress Raaen, A.M., Horsrud, P., Kjørholt, H., Økland, D., 2006. Improved routine estimation of the
can be used as the lower bound horizontal stress in drawing a stress minimum horizontal stress component from extended leak-off tests. Int. J. Rock
Mech. Min. Sci. 43, 37–48.
polygon. This improves the stress polygon which is lithology-depen- Saffer, D.M., Marone, C., 2003. Comparison of smectite- and illite-rich gouge frictional
dent. This improved stress polygon can narrow the area of the conven- properties: application to the updip limit of the seismogenic zone along subduction
tional stress polygon in certain conditions, e.g. in shales. A new method megathrusts. Earth Planet. Sci. Lett. 215, 219–235.
Schmitt, D.R., Currie, C.A., Zhang, L., 2012. Crustal stress determination from boreholes
of the minimum horizontal stress calculation is also proposed. In this
and rock cores: fundamental principles. Tectonophysics 570, 1–26.
method, the minimum horizontal stress is the one from the uniaxial Sibson, R.H., 1974. Frictional constraints on thrust, wrench and normal faults. Nature 247,
strain model plus a term which relates to Poisson's ratio and the mini- 542–544.
mum horizontal strain or vertical stress. Case applications demonstrate Takahashi, M., Mizoguchi, K., Kitamura, K., Masuda, K., 2007. Effects of clay content on the
frictional strength and fluid transport property of faults. J. Geophys. Res. 112, B08206.
that the proposed method provides better results in the horizontal Tembe, S., Lockner, D.A., Wong, T.F., 2010. Effect of clay content and mineralogy on fric-
stress estimates and in-stress constraints. tional sliding behavior of simulated gouges: binary and ternary mixtures of quartz, il-
lite, and montmorillonite. J. Geophys. Res. 115, B03416.
Townend, J., Zoback, M.D., 2004. Regional tectonic stress near the San Andreas fault in
central and southern California. Geophys. Res. Lett. 31 (L15S11).
References Warpinski, N.R., Teufel, L.W., 1989. In-situ stresses in low permeability, nonmarine rocks.
J. Pet. Technol. 41 (4), 405–414 (SPE-16402-PA).
Anderson, E.M., 1951. The Dynamics of Faulting and Dyke Formation with Application to
Warpinski, N.R., Branagan, P., Wilmer, R., 1985. In-situ Stress Measurements at U.S. DOE's
Britain. second ed. Oliver and Boyd, Edinburgh.
Multiwall Experiment Site. Mesaverde Group, Rifle, Colorado, pp. 527–536 (SPE
Barton, C.A., Zoback, M.D., Burns, K.L., 1988. In-situ stress orientation and magnitude at
12142, JPT).
the Fenton geothermal site, New Mexico, determined from wellbore breakouts.
White, A.J., Traugott, M.O., Swarbrick, R.E., 2002. The use of leak-off tests as means of
Geophys. Res. Lett. 15, 467–470.
predicting minimum in situ stress. Pet. Geosci. 8, 189–193.
Biot, M., 1941. General theory of three dimensional consolidation. J. Appl. Phys. 12,
Whitehead, W.S., Hunt, E.R., Finley, R.J., Holditch, S.A., 1986. In-Situ Stresses: a comparison
155–164.
between log-derived values and actual field-measured values in the Travis Peak for-
Bird, P., Kong, X., 1994. Computer simulations of California tectonics confirm very low
mation of East Texas. Paper SPE 15209.
strength of major faults. Bull. Geol. Soc. Am. 106, 159–174.
Whitehead, W.S., Hunt, E.R., Holditch, S.A., 1987. The effects of lithology and reservoir
Bishop, A.W., Green, G.E., Garga, V.K., Andresen, A., Brown, J.D., 1971. A new ring shear ap-
pressure on the in-situ stress in the Waskom (Travis peak) field. Paper SPE/DOE
paratus and its application to the measurement of residual shear strength.
16403.
Geotechnique 21, 273–328.
Wileveau, Y., Cornet, F.H., Desroches, J., Plumling, P., 2007. Complete in situ stress deter-
Bradley, W.B., 1979. Failure of inclined boreholes. Trans. ASME 101, 232–239.
mination in an argillite sedimentary formation. Phys. Chem. Earth 32, 866–878.
Breckels, I.M., van Eekelen, H.A.M., 1982. Relationship between horizontal stress and
Zhang, J., 2011. Pore pressure prediction from well logs: methods, modifications, and new
depth in sedimentary basins. SPE 10336. JPT 1982, 2191–2199.
approaches. Earth-Sci. Rev. 108, 50–63.
Bredehoeft, J.D., Wolff, R.G., Keys, W.S., Shuter, E., 1976. Hydraulic fracturing to determine
Zhang, J., 2013a. Borehole stability analysis accounting for anisotropies in drilling to weak
the regional in situ stress field in the Piceance Basin, Colorado. Geol. Soc. Am. Bull. 87,
bedding planes. Int. J. Rock Mech. Min. Sci. 60, 160–170.
250–258.
Zhang, J., 2013b. Effective stress, porosity, velocity and abnormal pore pressure prediction
Byerlee, J., 1978. Friction of rocks. Pure Appl. Geophys. 116, 615–626.
accounting for compaction disequilibrium and unloading. Mar. Pet. Geol. 45, 2–11.
Carena, S., Moder, C., 2009. The strength of faults in the crust in the western United States.
Zhang, J., Roegiers, J.-C., 2005. Double porosity finite element method for borehole model-
Earth Planet. Sci. Lett. 287, 373–384.
ing. Rock Mech. Rock. Eng. 38, 217–242.
Carpenter, B.M., Marone, C., Saffer, D.M., 2011. Weakness of the San Andreas Fault re-
Zhang, J., Roegiers, J.-C., 2010. Discussion on “integrating borehole-breakout dimensions,
vealed by samples from the active fault zone. Nat. Geosci. 4, 251–254.
strength criteria, and leak-off test results, to constrain the state of stress across the
Collettini, C., Niemeijer, A., Viti, C., Smith, S., Marone, C., 2011. Fault structure, frictional
Chelungpu Fault, Taiwan”. Tectonophysics 492, 295–298.
properties and mixed-mode fault slip behavior. Earth Planet. Sci. Lett. 311, 316–327.
Zhang, J., Bai, M., Roegiers, J.-C., 2003. Dual-porosity poroelastic analyses of wellbore sta-
Cui, J., et al., 2014. Determination of three-dimensional in situ stresses by anelastic strain
bility. Int. J. Rock Mech. Min. Sci. 40, 473–483.
recovery in Wenchuan Earthquake Fault Scientific Drilling Project Hole-1 (WFSD-1).
Zhang, J., Standifird, W., Lenamond, C., 2008. Casing ultradeep, ultralong salt sections in
Tectonophysics 619-620, 123–132.
deep water: a case study for failure diagnosis and risk mitigation in record-depth
Detournay, E., Cheng, A.H.-D., 1988. Poroelastic response of a borehole in a non-hydro-
well. Paper SPE 114273.
static stress field. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25, 171–182.
Zhang, J., Lang, J., Standifird, W., 2009. Stress, porosity, and failure dependent compres-
Eaton, B.A., 1975. The equation for geopressure prediction from well logs. Soc. Pet. Eng.
sional and shear velocity ratio and its application to wellbore stability. J. Pet. Sci.
AIME. Paper SPE 5544.
Eng. 69, 193–202.
Edwards, S.T., Meredith, P.G., Murrell, S., 1998. An investigation of leak-off test data for es-
Zoback, M.D., et al., 2003. Determination of stress orientation and magnitude in deep
timating in-situ stress magnitudes: application to a basin wide study in the North
wells. Int. J. Rock Mech. Min. Sci. 40, 1049–1076.
Sea. SPE/ISRM 47272.
Zoback, M., Hickman, S., Ellworth, W., 2010. Scientific drilling into the San Andreas fault
Engelder, T., Fischer, M.P., 1994. Influence of poroelastic behavior on the magnitude of
zone. Eos 92, 197–204.
minimum horizontal stress, Sh in overpressured parts of sedimentary basins. Geology
22, 949–952.

You might also like