JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. 0, 10.1029/2000JB000138, 2002
Published in Journal of Geophysical Research, vol. 107 (no. B2), cn: 2030,
doi: 10.1029/2000JB000138, pp. ESE.2.1 - ESE.2.17, February 2002.
Pore pressure and poroelasticity effects in Coulomb
stress analysis of earthquake interactions Errata Corrige page
attached; one item
Massimo Cocco refers to an error in the
Istituto Nazionale di Geofisica e Vulcanologia, Rome, Italy published version, not
present here.
James R. Rice
Engineering Sciences and Geophysics, Harvard University, Cambridge, Massachusetts, USA
Received 3 January 2001; revised 2 October 2001; accepted 7 October 2001; published XX Month 2002.
[1] Pore pressure changes are rigorously included in Coulomb stress calculations for fault
interaction studies. These are considered changes under undrained conditions for analyzing very
short term postseismic response. The assumption that pore pressure is proportional to fault-
normal stress leads to the widely used concept of an effective friction coefficient. We provide an
exact expression for undrained fault zone pore pressure changes to evaluate the validity of that
concept. A narrow fault zone is considered whose poroelastic parameters are different from those
in the surrounding medium, which is assumed to be elastically isotropic. We use conditions for
mechanical equilibrium of stress and geometric compatibility of strain to express the effective
normal stress change within the fault as a weighted linear combination of mean stress and fault-
normal stress changes in the surroundings. Pore pressure changes are determined by fault-normal
stress changes when the shear modulus within the fault zone is significantly smaller than in the
surroundings but by mean stress changes when the elastic mismatch is small. We also consider an
anisotropic fault zone, introducing a Skempton tensor for pore pressure changes. If the anisotropy
is extreme, such that fluid pressurization under constant stress would cause expansion only in the
fault-normal direction, then the effective friction coefficient concept applies exactly. We finally
consider moderately longer timescales than those for undrained response. A sufficiently
permeable fault may come to local pressure equilibrium with its surroundings even while that
surrounding region may still be undrained, leading to pore pressure change determined by mean
stress changes in those surroundings. INDEX TERMS: 7209 Seismology: Earthquake dynamics
and mechanics, 7260 Seismology: Theory and modeling, 7215 Seismology: Earthquake
parameters; KEYWORDS: Fault interaction, fluid flow, poroelasticity, effective friction, crustal
anisotropy
1. Introduction fault plane to compute Coulomb stress changes. In the framework
of the Coulomb criterion, failure on a fault occurs when the applied
[2] Earthquakes produce changes in the state of strain and stress increment, defined as
stress in the volume surrounding the causative faults. Coseismic
stress and strain changes caused by shear dislocations are
CFF ¼t þ mðs þ pÞ; ð1Þ
usually calculated using the numerical procedure proposed by
Okada [1985, 1992]. This approach is based on the solution of
the elastostatic equations in an elastic, isotropic homogeneous overcomes a stress threshold, where. t is the shear stress change
half-space. The coseismic strain and stress fields can be (computed in the slip direction), s .is the fault-normal stress
computed if the geometry and the slip distribution on the change (positive for extension), p .is the pore pressure change
rupturing fault plane are known (see Okada [1992], Stein et within the fault, and m is the friction coefficient which ranges
al. [1992], King et al. [1994], Stein [1999], and King and between 0.6 and 0.8 for most rocks [see Harris, 1998, and
Cocco [2000], among several others). In the near field the references therein]. The quantities included in (1) should be
induced coseismic stress consists both of a dynamic (transient) considered as functions of time.
and a static (permanent) perturbation. The calculation of [4] Earthquakes perturb the state of stress of a crustal volume,
dynamic stress changes requires the solution of the elastody- and they cause a variety of hydrologic phenomena [Scholz, 1990;
namic equations. It implies that both shear and fault-normal Sibson, 1994; King and Muir-Wood, 1994; Roeloffs, 1996, 1998;
stresses can vary as functions of times reaching the static Roeloffs and Quilty, 1997]. Some of these effects can be explained
configuration after a few tens of seconds [Harris and Day, by the poroelastic response to the earthquake-induced strain field.
1993; Cotton and Coutant, 1997; Belardinelli et al., 1999, and Pore pressure changes modify the coseismic stress redistribution,
references therein]. and for this reason they are included in the definition of the
[3] Fault interaction is currently investigated by means of these Coulomb failure function (1). Because the coseismic stress changes
analytical formulations and using the induced stress on a specified occur on a timescale that is too short to allow the loss or gain of
pore fluid by diffusive transport (fluid flow), the pore pressure
changes included in (1) on that timescale are associated with the
Copyright 2002 by the American Geophysical Union. undrained response of the medium [Rice and Cleary, 1976]. Later,
0148-0227/02/2000JB000138$09.00 we discuss somewhat longer timescales for which the fault is no
ESE X-1
ESE X-2 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
longer undrained. The undrained conditions are those for which pressure changes are time-dependent; therefore it is necessary to
there is no fluid flow. From an analytical point of view the specify the timescale during which the poroelastic model is
undrained response implies that the fluid mass content per unit applied.
volume is constant (m = 0), but the pore pressure is altered. [8] This paper discusses the assumptions required to correctly
Under these conditions the relationship between stress and strain include the pore pressure changes in Coulomb stress modeling
for a fluid-infiltrated poroelastic material is equivalent to an and provides a more general expression for the effective normal
ordinary elastic material with appropriate coefficients for the stress for different timescales. We start investigating the short-
undrained conditions [Rice and Cleary, 1976; Roeloffs, 1996]. term postseismic period, in which both the fault zone and the
[5] According to Rice and Cleary [1976] the pore pressure adjoining lithosphere respond under undrained conditions. In this
change resulting from a change in stress under undrained con- case we first assume that the fault zone is a poroelastic isotropic
ditions is given by medium, but we will also consider the effect of anisotropy
within the fault zone. Then we study an intermediate timescale,
skk which will exist for a sufficiently permeable fault, during which
p ¼ B ; ð2Þ the fault core reaches a local pressure equilibrium with its
3
lithospheric surroundings, while the adjoining lithosphere is still
responding as if it were undrained. We will not consider here
where B is the Skempton coefficient [Skempton, 1954; Kuempel, longer timescales during which the transition from short-term
1991]. Rice and Cleary [1976], Roeloffs and Rudnicki [1985], and undrained response to long-term drained response takes place
Roeloffs [1996] present a compilation of experimental determina- also in the surrounding lithosphere.
tions of B indicating a range between 0.5 and 0.9. In Coulomb
stress analysis [see Stein et al., 1992; Harris and Simpson, 1992;
King et al., 1994; Harris, 1998, and references therein] it is
assumed that for plausible fault zone rheologies the change in pore 2. Poroelastic Constitutive Relations
pressure becomes proportional to the fault-normal stress: [9] The stress-strain relation for an ordinary isotropic linearly
elastic solid can be expressed as
^
p ¼ Bs: ð3Þ
l
2Geij ¼ sij skk dij ; ð5Þ
This is certainly true if in the fault zone s11 = s22 = s33, so 3l þ 2G
that skk/3 = s and (2) becomes (3) [see Simpson and
Reasenberg, 1994; Harris, 1998]. By substituting (3) in (1), we where eij and sij are the strain and stress tensors, respectively; G
obtain and l are the Lamé parameters (G is the rigidity) and dij is the
Kronecker delta. Hooke’s law (5) can be rewritten using the
CFF ¼tþm0 s; ð4Þ Poisson ratio v.. as
v
where m0 = m(1B̂) is the effective (or apparent) friction coefficient. 2Geij ¼ sij skk dij : ð6Þ
Equation (4) is very common in the literature, and it has been 1þv
widely used to calculate Coulomb stress changes [see Harris,
1998, and references therein]. A variety of values are used for these Because here we consider linear elasticity, these constitutive
calculations: the friction coefficient m ranges between 0.6 and 0.8, relations must be applied to small stress-strain magnitudes. We
while B ranges between 0.5 and 1 [Green and Wang, 1986; Hart, assume that they are valid in such isotropic form for coseismic
1994]. The resulting values for the effective friction coefficient stress-strain changes caused by shear dislocations, which are of
range between 0.0 and 0.75 (0.4 has been used in many interest since we do not know the absolute value of the regional
calculations by Stein et al. [1992] and King et al. [1994]). Several remote tectonic stress.
studies have concluded that Coulomb stress modeling is only [10] Because compact rocks consisting of solid phase materials
modestly dependent on the assumed value of the effective friction are not an appropriate model for the crust, we have to consider our
coefficient [see King et al., 1994]. This result might depend on the medium as porous or cracked. The stress-strain relations for a
choice of the poroelastic model (equation (2) or (3)) in Coulomb poroelastic medium are slightly different from (6) because they
stress analyses. It is important to emphasize that the effective include the pore pressure term [Biot, 1941, 1956; Rice and Cleary,
friction coefficient m’ is not a material property, but it depends on 1976]. According to Rice and Cleary [1976], these constitutive
the ratios of stress changes in the medium [Byerlee, 1992; Hill et relations are
al., 1993; Beeler et al., 2000].
[6] Several recent papers have focused attention on the corre- v 3ðvu vÞ
lation between fault-normal stress changes and earthquake loca- 2Geij ¼ sij skk dij þ pdij ð7aÞ
1þv Bð1 þ vÞð1 þ vu Þ
tions as well as seismicity rate changes [see Perfettini et al., 1999;
Parsons et al., 1999; Cocco et al., 2000, and references therein].
However, it is still not well understood why these fault-normal 3r0 ðvu vÞ 3
m ¼ m0 þ skk þ p ; ð7bÞ
stress changes should provide a better explanation of this correla- 2GBð1 þ vÞð1 þ vu Þ B
tion than Coulomb stresses. Fluid flow (time-dependent) as well as
the choice of the proper expression for p in Coulomb analyses where m0, r0 are the fluid mass content and the density measured
might help to explain some aspects of this paradox. with respect to a reference state at which we take P = 0. Here v is
[7] Beeler et al. [2000] pointed out that using the constant the Poisson ratio under drained conditions, whereas the term vu
apparent friction model (equation (4)) in Coulomb analyses may represents the undrained Poisson ratio, which is a function of v, the
provide a misleading view in estimating stress changes. They bulk modulus (K), and the Skempton coefficient (B) of the medium
compare that model with an isotropic and homogeneous (same [see Rice and Cleary, 1976; Kuempel, 1991]. Equation (7b) shows
properties within the fault as outside) poroelastic model, equations that for undrained conditions (m = 0), p = Bskk/3 yielding
(1) and (2), and conclude that Coulomb failure stress shows (2) when pore pressure and mean stress changes are considered.
considerable differences for different tectonic environments. It is Equation (7a) is equivalent to (6) for a poroelastic medium if v in
important to emphasize that because of fluid flow the induced pore (6) is replaced by vu. In fact, using the relation (2) in (7a) to
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X-3
Surrounding Crust strain components are the same inside and outside the fault zone
[Rice, 1992], namely, e110 = e11, e220 = e22, e120 = e12. Similarly,
G, λu,Ku, νu conditions of mechanical equilibrium require that certain stress
3
1 components must be the same everywhere within the fault zone as
in the nearby crust outside it, namely, s330 = s33, s310 = s31, s320 =
s32. In other words, the equilibrium conditions on the problem
h
2 restrict s33 .to continuity but leave the other normal components
unrestricted. This means that the other stress components inside the
fault zone may be different from those outside it. They are
determined from mechanical constitutive relations.
homogeneous [14] According to (5) we have the following relations:
isotropic Fault Zone Parameters
medium G', λ'u,K'u, ν'u 1
2lu
2ðe11 þ e22 Þ ¼ s11 þ s22 skk ð10aÞ
G 3lu þ 2G
Figure 1. Undrained poroelastic fault model.
1 2l0
2 e011 þ e022 ¼ 0 s011 þ s022 0 u 0 s0kk ; ð10bÞ
describe undrained (e.g., coseismic) strain and stress changes, we G 3lu þ 2G
get the following constitutive relation:
where primes denote quantities inside the fault zone. Using the
vu continuity conditions for the strain components appearing in (10a)
2Geij ¼ sij skk dij : ð8Þ
1 þ vu and (10b) and for the fault-normal stress, we can write
Equation (8) is equivalent to (6) but now represents a poroelastic 1 lu þ 2G 1 l0u þ 2G 0
medium under undrained conditions. skk s33 ¼ 0 s s33 : ð11Þ
G 3lu þ 2G G 3l0u þ 2G0 kk
[11] The following relation [Rice and Cleary, 1976] relates the
undrained Poisson ratio to the other poroelastic parameters:
[15] For a slightly more concise notation, let M = l + 2G be the
modulus for one-dimensional strain and recall that K = l + 2G/3.
3v þ Bð1 2vÞ 1 KKs Then (11) becomes
vu ¼ ; ð9Þ
3 Bð1 2vÞ 1 KKs
1 Mu skk 1 Mu0 s0kk
s33 ¼ 0 s 33 ;
G Ku 3 G Ku0 3
where K is the bulk modulus of the saturated rock under drained
conditions and Ks is a modulus which, for certain simple materials where the quantity Mu/Ku corresponds to (1 vu)/(1 + vu), which
(uniform properties of solid phase in response to hydrostatic might be alternatively used in the following equations (with primes
stressing, fully interconnected pore space), can be equated to the denoting the values within the fault zone). Solving the previous
bulk modulus of the solid grains in the rock. We emphasize that if relation for skk0/3 within the fault zone, we get
K = Ks, then vu = v. In general, the undrained bulk modulus is
larger than the drained one, and Ks > Ku > K. According to (9) the
s0kk Ku0 G0 Mu skk G G 0
undrained Poisson ratio is larger than the drained Poisson ratio (vu ¼ 0 þ s33 : ð12aÞ
> v). The same is true for the Lamé parameter, lu > l. The rigidity, 3 Mu G Ku 3 G
on the contrary, remains the same, Gu = G. These considerations
suggest that the fault stiffness for undrained conditions is larger To emphasize that this applies for the stress changes caused by a
than that for drained ones. nearby earthquake, we write
3. Pore Pressure Changes in an Undrained s0kk K 0 G0 Mu skk G G 0
¼ u0 þ s33 ð12bÞ
Poroelastic Fault Model 3 Mu G Ku 3 G
[12] We investigate the stress conditions for a fault in a and understand the unprimed stress changes (sij) to be those
poroelastic medium. The fault zone materials have different conventionally computed by elastic dislocation theory. For
properties with respect to the surroundings. Let 1 and 2 represent Coulomb analysis we need to know the pore pressure changes
the coordinate directions in the fault zone and 3 represent the induced in the fault zone, which we obtain by substituting (12b)
coordinate direction perpendicular to the fault plane (Figure 1). in (2):
The stress-strain relations for the medium are given by (8). We
0
interpret the elastic moduli here as moduli for undrained defor- 0
0
0 skk
0
0 Ku G Mu skk G G0
mation. We indicate with G0, lu0 , Ku0 , vu0 the Lamé and bulk p ¼ B ¼ B 0 þ s33 ; ð13Þ
3 Mu G Ku 3 G
moduli and the Poisson ratio within the fault zone, while G, lu,
Ku, vu denote the parameters in the surrounding crust. In the where B0 is the Skempton coefficient in that fault zone. Equation
following, we use the stress-strain relation for a poroelastic (13) shows that induced pore pressure changes depend both on the
medium in the form of (5), using the Lamé moduli. mean stress and the fault-normal stress changes. The relevant
[13] Considering the conditions of mechanical equilibrium and effective normal stress change for Coulomb stress analysis is thus
strain compatibility, there exist equality conditions for strain and
stress components within the fault zone and outside it, in the cases 0
0 0 0 skk
that we consider here, for which fault zone thickness is much less seff
33 ¼ s33 þ p ¼ s33 B
than length scales over which stress and strain vary outside the 3
K 0 G0 Mu skk G G0
fault. We indicate with eij0 and sij0 .the strain and the stress tensors ¼ s33 B0 u0 þ s33 : ð14Þ
within the fault zone. Kinematic compatibility implies that certain Mu G Ku 3 G
ESE X-4 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
[16] Equations (13) and (14) give the relative weights of fault- [20] If the density contrast between the fault zone materials and
normal and mean stress perturbations in determining the pore the surrounding crust is not very large (r r0), the relative
pressure and Coulomb stress changes. In general, the pore pressure variation of G and M can be expressed as
changes depend on both these quantities. In particular, the pore
pressure changes are related to the fault-normal stress changes G G 0 rVS2 r0 VS0 2 VS 2 VS0 2
through the rigidity contrast between fault zone materials and the ¼ ;
surrounding crust. G rVS 2 VS2
0 2
[17] We can recognize two limiting cases for (14). The first one G0 VS
holds when G0 = G and the mean stress is the relevant quantity. ¼ ; ð19Þ
G VS
Thus (14) becomes 0 2
Mu0 VP
¼ :
0 Ku0 Mu skk Mu VP
seff
33 ¼ s33 B ; ð15Þ
Ku Mu0 3
[21] An opposite limiting case exists when the rigidity contrast
the second limiting case is obtained when G G0, and (14) is negligible (G0 approximately equal to G), and therefore we have
becomes
0
VS2 VS0 2 r0 r
0 Ku :
seff
33 ¼ 1 B s33 : ð16Þ VS2 r0
Mu0
[18] Equations (15) and (16) express the effective normal stress [22] This latter case represents the limiting case (G0 approxi-
changes for the case where pore pressure changes depend solely on mately equal to G) yielding (15), and mean stress perturbations are
mean or fault-normal stress changes, respectively. Equation (16) is the only contribution to the effective normal stress changes.
equivalent to the effective friction approach of (4) if the Skempton [23] In a first simplified model we assume that the fault zone is a
parameter is given by B̂ = B0K0u/M0u. solid of the same lithology as the adjoining crust, densely fractured
with an isotropic distribution of cracks saturated by fluids, while
the crust is considered as a much less cracked but still saturated
4. Elastic Moduli in the Fault Zone Poissonian body (l = G). We consider that none of the crack walls
can open or close toward one another when an isotropic stress is
[19] Equation (13) relates pore pressure changes to fault-
applied, so they produce no change in volume and hence no
normal and mean stress changes through two factors, which
alteration of K0u. We can thus observe that the bulk modulus K0u
depend on the elastic parameters in the fault zone and in the
is unaffected by the presence of saturated cracks, at least assuming
surrounding crust. The variation of these elastic parameters is
that their aspect ratio is much less than the ratio of liquid bulk
reflected in the variation of P and S wave velocities. There-
modulus to solid bulk modulus, as pointed out by O’Connell and
fore information on shear wave velocity anomalies in the fault
Budiansky [1974]. Assuming that the fault zone has the same
zone might be used to constrain numerical values of the
lithology as the surroundings, just much more cracked, implies that
factors appearing in (13) and to discuss the two limiting
K0u = Ku. This also means that r = r0, neglecting the crack space
cases reported in (15) and (16). In particular, the following
contributions to volume. In these conditions the crack walls can
relations hold (assuming that measured P wave speeds corre-
slide in shear, so that G is reduced (G0 < G), yielding (19).
spond approximately to undrained response in the sense of
[24] We use seismic evidence for P and S wave velocity
poroelasticity):
variations to infer possible values of the elastic moduli in the fault
2 2 zone. Studies of local crustal tomography provide evidence on the
Mu VP Mu V 4 body wave velocity variations in fault zones. Although several
¼ ; Ku ¼ G 43 ¼ G P2 ; ð17Þ studies (mostly based on VP tomographic images) interpreted the
G VS G VS 3
fault zone as a high-velocity body [Lees, 1990; Lees and Nich-
olson, 1993; Zhao and Kanamori, 1993, 1995], many others have
and the same relations with primes (K0u, M0u) indicate the moduli
suggested the presence of fluids within the fault zone [Eberhart-
inside the fault zone. This equation yields
Phillips and Michael, 1993; Johnson and McEvilly, 1995; Thurber
et al., 1997]. Zhao et al. [1996] and Zhao and Negishi [1998]
Ku 4 VS 2 found evidence of low P and S wave velocities and high Poisson
¼1 ð18Þ
Mu 3 VP ratio at the hypocenter of the 1995 Kobe earthquake. We remark
that VS and the Poisson ratio. vu (or VP/VS) are much more
for the quantity which appears in (14). The square of the S wave sensitive to fluids than VP [see also Eberhart-Phillips and Reyners,
velocity anomaly of the fault zone with respect to the 1999].
surrounding crust is related to the density and rigidity of the [25] Studies on fault zone trapped waves yield more useful
two media: constraints to the quantities defined in (17) and (19) because they
have an optimal resolution of the inner structure of fault zones
whose thickness can range between 20 and 400 m [Li et al., 1990,
VS2 VS0 2 r0 G rG 0 1994]. Li and Leary [1990] have shown the fracture density and the
¼ ;
VS2 r0 G S wave velocity model for the Oroville (California) fault zone as
determined by body wave travel time modeling. They clearly show
where r is the density. A similar relation holds for the that the fault zone corresponds to a reduction in shear wave
modulus for one-dimensional strain M and the P wave velocity velocity (roughly 50%) and an increase of the crack density (up
anomaly: to 0.75). Li et al. [1990, 1994] point out that the reduction in S
wave velocity inside the fault zone ranges from 30 to 50%.
According to Mooney and Ginzburg [1986] and Li and Leary
VP2 VP0 2 r0 Mu rMu0
¼ : [1990] the velocity structure of the Calaveras fault shows a P wave
VP2 r0 Mu velocity reduction of nearly 30%.
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X-5
[26] Although our review of velocity models of fault zones is far estimated does not affect the relative importance of mean stress
from complete, we use these representative values to estimate the and fault normal stress in (13) for p0; it only affects the factor K0u/
quantities defined above. According to the aforementioned studies Mu0 in front.
and to relations (17), (18), and (19) if V S0 /VS ranges between 0.5
and 0.7, the rigidity ratio G’/G ranges between 0.25 and 0.5. This
implies that the rigidity reduction (G G0)/G is between 0.75 and 5. Modeling Static Stress Changes
0.5. As expected, if the reduction in S wave velocity within the From Shear Dislocations
fault zone is very large (50 – 70%) the ratio G0/G becomes much
smaller than (G G0)/G. Moreover, if we assume that the reduction [32] In this section we aim to compare the shear, fault-
in S wave velocity is larger than that in P wave, then the V P0 /V S0 normal, and mean static stress changes caused either by a
ratio inside the fault zone is larger than the corresponding value in vertical strike-slip fault or a normal fault in an elastic homoge-
the surrounding crust. We consider the simplified model described neous half-space. We use the three-dimensional (3-D) dislocation
above, which yields K u0 = Ku (= 5G/3 in a Poissonian surrounding code developed by Nostro et al. [1997], which is based on
crust), and then given G0/G, we can calculate M u0 : numerical representation provided by Okada [1985, 1992].
Again, our discussion here is directed to the short timescale
for which the fault and its surroundings respond as if undrained.
4 4 G0 5 4 G0
Mu0 ¼ Ku0 þ G0 ¼ K u þ G¼ þ G: The elastic stress changes caused by shear dislocations illustrate
3 3 G 3 3 G
the spatial variability and absolute values of the different terms
in (1), (2), and (4) and allow a comparison between the
[27] This results in Coulomb stress changes resulting from the application of (1)
as opposed to (4). Figure 2a shows the shear, normal, and
Ku0 Ku 5 G 5 Coulomb stress changes caused by a vertical strike-slip fault
¼ ¼ ¼ 0 :
Mu0 Mu0 3 Mu0 5 þ 4 GG mapped both on a horizontal layer at 6 km depth as well as on
a vertical cross section A-A’ perpendicular to the fault strike.
Coulomb stress changes have been computed by means of (4)
[28] According to this relation the ratio Ku0 /M u0 ranges between
on secondary faults having the same orientation and mechanism
0.714 and 0.883 for G0/G between 0.5 and 0.25, while it is equal to
of the causative fault and using a constant effective friction
0.556 for G0/G = 1. As an example, we provide a tentative estimate
equal to 0.4 (corresponding to m = 0.75 and B = 0.47). King et
of the two proportionality factors that appear in (13). We assume a
al. [1994] have already discussed these stress patterns in detail;
reduction in P and S wave velocities in the fault zone of 18% and
here we only remark that positive stress changes can favor
50%, respectively. These assumptions yield
failures on appropriately oriented planes. We also point out that
the variability of the stress patterns on the horizontal maps,
G G0 G0 Ku K0 where both s1 and s3 lie, is more evident for strike-slip faults
¼ 0:75; ¼ 0:25; ¼ 0:556; u0 ¼ 0:883:
G G Mu Mu than on the vertical cross sections. Moreover, the amplitudes of
fault-normal stress changes at depth are quite small. It is
[29] Using these values, the constants which multiply the mean important to remark that the cross section is not taken in the
and the fault-normal stress changes in (13) and (14) are 0.45 and middle of the rupturing fault because such a direction is nodal
0.75, respectively. The latter does not seem to be negligible at all, for fault-normal stress changes, as shown in the map view.
and for this illustration, Figure 2b shows similar results for a normal fault dipping 70 to
the east; we only show the E-W vertical cross section calculated
in the middle of the fault. As expected for a normal fault, the
skk
p0 ¼ B0 0:375 þ 0:625s33 largest spatial variability of stress changes occurs on the vertical
3 plane used for the sections where both s1 and s3 lie.
¼ B0 ½0:125ðs11 þ s22 Þ þ 0:750s33
: [33] Figure 3 shows the mean stress changes caused by the
vertical strike-slip fault as well as the stress changes along the fault
[30] Thus the assumption that p0 is proportional only to skk/ plane directions (1 and 2 in Figure 1). The stress changes in the
3 does not seem to be strongly supported by observation of P and S direction perpendicular to the fault plane (fault-normal stress) are
wave velocities, but at the same time, neglecting this term may be shown in Figure 2a (middle). It emerges from these calculations
justifiable only if the reduction in S wave velocity is much larger that the three diagonal terms of coseismic stress changes (sii) are
than 50%, at least under the conditions assumed in the simplified substantially different in amplitude. In particular, the largest
model considered here. This model can be reasonable for faults that amplitudes are found for the component oriented along the slip
have experienced little slip, but it might be not reliable for a direction and that perpendicular to the fault plane, respectively.
relatively mature fault zone. In this latter case, the presence of a Similar results have been obtained for a normal fault. The stress
fault gouge with a different porosity, and possibly fluid-altered change is maximum in the direction of slip; this means s11 for a
composition, with respect to the host lithology might be more strike slip fault. This result is also evident at depth, as shown in the
properly represented by a density contrast [Mooney and Ginzburg, vertical cross sections in Figure 3. This implies that the condition
1986], so that r0 < r. Even in this case, we can show that the change s11 = s22 = s33, which leads to (4), is not satisfied in the
in the one-dimensional strain modulus M is larger than the change volume surrounding the fault; moreover, the spatial variations of
in density. According to Mooney and Ginzburg [1986] we can these stress components are quite different. This observation
assume that VP/VP = Fr/r, with F 1, and write further supports the conclusion that (3) must be justified by
different considerations, rather than assuming s11 = s22 =
s33.
VP 1 Mu0 r Mu0 r
¼ ¼ ð1 þ 2F Þ : [34] In Figure 4 we show the Coulomb stress changes calculated
VP 2 Mu0 r Mu0 r both by (4) (the constant effective friction model) and by (1) and
(2) (the isotropic model, see Beeler et al. [2000]). The two maps on
[31] This gives a greater change in modulus M0u rather than in r the top of Figure 4 represent the Coulomb stress changes computed
so that, approximately, neglect of r .changes in our estimate of using the same values of friction and Skempton parameters (0.75
modulus changes, as in (19), is still valid. Note that for a given G0/ and 0.47, respectively). In Figure 5 we show a similar comparison
G, a different reduction from Mu to M0u than what we have for a normal fault in vertical cross section. It emerges from these
ESE X-6 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
Figure 2. Shear, normal, and Coulomb stress changes caused by (a) a vertical strike-slip fault and (b) a 70 dipping
normal fault calculated using the constant effective friction model of equation (4). The amount of slip on the rupturing
fault is 50 cm. Coulomb stress changes have been computed on secondary fault planes having the same geometry and
mechanisms as the causative faults. For all these calculations, m0 = 0.4, which corresponds to m = 0.75 and B = 0.47.
The vertical cross section for the normal fault case shown in Figure 2b is taken in the middle of the causative fault.
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X-7
Figure 3. Map view and vertical cross section of mean stress changes (skk/3) and induced stress perturbations for
the two isotropic components oriented along the fault directions (1 and 2 of Figure 1) caused by a vertical strike-slip
fault as shown in Figure 2a.
ESE X-8 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
Figure 4. Coulomb stress changes at 6 km depth caused by a vertical strike-slip fault computed with the constant
apparent friction model (equation (4)) and the isotropic friction model (equations (1) and (2)). For this latter model we
show the calculations using different values of the friction and Skempton coefficients.
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
ESE
X-9
Figure 5. Same calculations as Figure 4 shown in a vertical cross section (as in Figure 2b) for a 70 dipping normal fault.
ESE X - 10 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
calculations that the equation adopted for computing Coulomb
stress affects the resulting spatial patterns. Such a result has been
already discussed by Beeler et al. [2000], who also quantified the
amount of such variations. They concluded that these two pore
pressure models yield considerable differences in the calculated
Coulomb stress changes for reverse, normal, and strike-slip faults.
They also suggest that the use of the constant effective friction
model (equation (4)) could lead to errors in estimating coseismic
stress changes. The calculated Coulomb stress changes also depend
on the assumed value of the Skempton parameter B [see also
Beeler et al., 2000]. In Figures 4 and 5 we show the results of
calculations using three different values of B between 0.2 and 1. As
expected, increasing B increases the Coulomb stress changes in the
off-fault lobes. This emphasizes the role of pore fluid pressure in
earthquake failure [see, e.g., Segall and Rice, 1995] but also brings
up the question of which is the most appropriate way to represent
coseismic pore pressure changes. The differences among the
Coulomb stress changes shown in Figures 4 and 5 can be as large
as several bars.
[35] To summarize, we have calculated the ratio between the
mean stress (changed in sign: skk/3) and the fault-normal stress
changes s(= s33)(see Figure 6a). Figure 6a shows that there is
a region along the fault strike direction where this ratio is highly
variable. Outside this region, on the two opposite sides of the fault,
this ratio is negative and smaller than unity: This means that the
mean stress and the fault-normal stress changes have the same sign
but the latter is larger than the former. On the contrary, in the area
where the ratio is positive they have opposite sign. A unitary value
for this ratio would imply that p0 = B0s. The strike direction
and that perpendicular to it are nodal for both these stress changes;
thus, in these zones the amplitudes of both mean stress and fault-
normal stress changes are very small (see Figures 2a and 3). In the
off-fault lobes, where both mean and fault-normal stress change
amplitudes are relevant, this ratio is negative and smaller than
unity. This is quite evident in Figure 6b, where we plot the
difference between normal and mean stress changes. These two
terms differ mostly at the ends of the slipped zone. The implication
of these calculations is that the effective friction coefficient is not
constant in the volume surrounding the causative fault. This is
expected, since from its definition it is a function of ratios of all the
fault-normal stress changes to one another in the medium.
[36] The considerations discussed above point out that in a
poroelastic isotropic medium the relation used to compute pore
pressure changes affects the calculation of Coulomb stress changes.
Effective normal stress depends on both mean stress and fault-
normal stress changes, and the proportionality factors do not
vanish, except in very special situations that might not be realistic
for actual fault zones. Moreover, the amplitudes of these stress
changes show different spatial patterns, depending on the faulting
mechanism [see also Beeler et al., 2000]. Therefore, in order to
choose the appropriate equation to compute Coulomb stress it is
necessary to consider more complex situations and different
properties for the fault zone materials.
6. Effects of Anisotropy Within the Fault Zone
[37] The results discussed above have been obtained under the
assumption that the fault zone materials are isotropic, i.e., that they
are permeated by an isotropic distribution of cracks that are
saturated by fluids. However, anisotropy within the fault zone
caused by aligned fractures may lead to different conclusions
concerning the proportionality between pore pressure and fault-
Figure 6. (a) Spatial pattern of the ratio between mean stress normal stress changes. Here we examine the effect of an aniso-
perturbations (changed in sign: skk/3) and fault-normal stress tropic distribution of cracks within the fault zone, and we derive an
changes (s = s33) for a vertical strike slip fault. (b) Spatial alternative formulation for the pore pressure changes for undrained
pattern of the difference between normal and mean stress changes deformation. Because we are interested here in the undrained
(skk/3). Worthy of note is the amplitude of such a difference. response of the medium, we do not discuss the anisotropy of
permeability of the fault zone materials.
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X - 11
[38] In the general case, possibly anisotropic, equation (2), Hence, for that type of anisotropy of pore space, which may be
which introduces the Skempton coefficient, must be generalized appropriate for a fault zone, it is correct to use the simplified
to the statement that a pore pressure change concept that induced pore pressure under undrained conditions is
determined solely by the change in fault-normal stress. In other
sij words, if the anisotropy of the fault zone material is so extreme that
p ¼ Bij ð20Þ when extracted from the fault and held at constant stress while fluid
3
is pumped into it, it expands only in the 3 direction, then p0 would
is induced by application of stress changes sij under undrained be determined solely by s033 = s33, and then the m0 concept would
conditions. The set of coefficients Bij constitute what we propose to apply exactly with B̂; = B033/3 in equation (3).
call a Skempton tensor. They reduce, of course, to Bij = Bdij in the [43] It might be interesting to know how much anisotropy is
isotropic case. If we regard P as a function of the set of stresses [s] needed to justify the effective friction concept in the way just
and of the fluid mass m, per unit volume of reference state, discussed. Recent papers on fault zone trapped waves [Leary et al.,
contained in the porous material, that is, p = p([s], m), then Bij 1987; Zhao and Mizuno, 1999] have shown very good quality data,
satisfy but unfortunately, there is still no answer to this question. Zhao and
Mizuno [1999] found a crack density distribution for the 1995
Kobe (Japan) earthquake that is smaller (0.2) than that expected for
Bij ¼ 3 @pð½s; m
Þ=@sij : ð21Þ a fracture zone (0.6 – 0.75). It is important to point out that the
presence of anisotropy within the fault zone may justify variations
[39] We show in Appendix A that an alternative, and instructive, of shear wave velocity of 50% and larger. Future observations are
interpretation can be obtained for that partial derivative once we needed to shed light on this problem; they will be helpful to
recognize [Rice and Cleary, 1976] that sijdeij + pd(m/r) must be a reconstruct the inner structure and mechanical properties of fault
perfect differential, where r is the density of the pore fluid zone materials.
(conceptually in a reservoir of pure fluid at local equilibrium with
the porous medium). Here it may be noted that m/r is the fluid
volume fraction (fluid volume per unit of reference state volume of 7. Short Time Pore Pressure Equilibrium Between
the porous material) in the case considered, when all pore space is Fault Zone and Adjoining Rock Mass
connected and fluid-infiltrated. The mass of fluid per unit reference
[44] We have focused thus far on the postseismic period, in
state volume is m and r .is the density of pore fluid at pressure p,
which the fault zone behaves as if it is undrained. However, if the
and we assume r = r( p). The differential form sums the work of
fault zone is moderately thin and has some permeability, then it is
stresses in moving the boundaries of an element of the porous
reasonable to expect that on what is also a relatively short time-
material and the work of pore pressure in enlarging the boundaries
scale, the fault zone will act as if it were locally drained and reach
of the pore space. Together the terms constitute the change dU in
pressure equilibrium with its surroundings, so that p0 evolves
the strain energy U of the solid phase, which must be a function of
toward p. p itself will be time-dependent because of the pore
state. As developed in Appendix A, this is equivalent to the
fluid fluxes set up by the gradients in the coseismically induced
familiar notion from the thermodynamics of mixtures that sijdeij
pore pressure field, but unless the fault zone is very thick and/or is
+ m̂dm sijdeij + m̂dm is a perfect differential, where m̂ is the
very impermeable compared to its surroundings, that variation of P
chemical potential of the pore fluid.
in the adjoining rock will have a much longer timescale than for
[40] By either route, the existence of the perfect differentials
local drained response of the fault. In such cases it is reasonable to
implies a Maxwell reciprocal relationship, which is shown in
expect that p0 will have relaxed to p well before p itself has
Appendix A to give the alternative interpretation of Bij as
relaxed much from its undrained value just after the earthquake
stress change. So, on such short but not extremely short timescale
Bij ¼ 3r @eij ð½s; m
Þ=@m : ð22Þ in which the fault acts as drained, but its surroundings remain
undrained, we get p0 p and thus p0 is proportional to the
The derivative corresponds to the change as fluid mass is pumped mean stress change outside the fault zone, p0 skk/3.
into the porous material under conditions for which all of the [45] In order to model this behavior in a simple way, we
stresses skl are held fixed. consider a fault zone of thickness h (see Figure 1) having a
[41] Thus, for example, in the isotropic case B/(3r) is the uniform permeability k0, fluid viscosity h0, and storage modulus
increase of each extensional strain per unit of increase dm of fluid N0. The surrounding crust is modeled as a pair of semi-infinite
mass pumped into the material, under conditions for which the total domains with corresponding parameters N, k, h. The storage
stresses are held constant. Equivalently, B/3 is the increase of each modulus is just the inverse of the storage coefficient (called S by
extensional strain per unit increase dm/r of fluid volume pumped in Wang [2000]), in response to pore pressure changes, for one-
under constant stresses. dimensional straining under constant stress in the straining
[42] Let suppose that within the fault zone there is a highly direction. See Appendix B for its precise definition and expres-
anisotropic distribution of cracks or flattened pores, lying so that sion in terms of moduli already introduced. As shown in Figure
their long directions are approximately parallel to the fault plane. 1, the axis 3 is perpendicular to the fault. The diffusivity inside
We isolate a sample of material of the fault and subject it to its in the fault zone (c0) and in the surrounding medium (c) can be
situ stress state s0kl. Then, holding those stresses constant, we respectively defined as
pump an increment dm0 of fluid mass into the porous material. In
that case, because of the assumed orientations of the flattened c0 ¼ k 0 N 0 =h c ¼ kN =h;
pores we would expect the fault-parallel strain increments de011
and de022 to be much smaller than the fault-normal component which depend on the viscosity, the permeability, and the storage
de033 because it is the latter which would be primarily influenced modulus (the increase of fluid mass content when the pressure
by fluid injection into the fault parallel-crack and pore space. varies under one-dimensional strain conditions) of the fault and of
That means B033 is much larger than the other components of B0ij, the surrounding crust.
and therefore that [46] By solving the one-dimensional consolidation problem for
a layer of one porous medium within an effectively infinite outer
ds0ij ds0 ds33 one, we model the evolution of the pore pressure within the fault
dp0 ¼ B0ij B033 33 ¼ B033 : ð23Þ
3 3 3 zone toward its longer time limit p. The pore pressure inside the
ESE X - 12 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
fault zone (p0) is a function of time and position in the fault zone, A) pore pressure evolution
but for simplicity, we solve the problem at the center of the fault short timescale
2 2
zone (Appendix B). However, in the following, we will derive an ∆ p ′(t) − ∆p td =
h
+
h N 2
analytical expression for the pore pressure at any point inside the +
∆ p ′(0 ) − ∆p 8 c ′ 8c N ′
fault zone. Our analysis of short-time undrained response shows
that at time t = 0+ the fault zone has the pore pressure change p0 1
(given by equation (13)), whereas the surrounding crust has the c′ N 2
r=
change p = Bskk/3. We determine the subsequent temporal c N′
evolution by finding the difference between the pore pressure 0.8
inside and outside the fault zone, normalized by its value just after
the time of the main shock (t = 0+): 0.6
r =1 r =∞
0 4
P~ ¼
p ðt Þ p
: ð24Þ 0.4
∞
p0 ð0þ Þ p
r =0
0 0 0.2 0.01
[47] Here the notation p (t) denotes p (z, t)jz=0,i.e.,the pore 0.11
pressure change at the center of the fault zone. In AppendixB 1
we solve simple diffusion equations using the Laplace transform 0
and its Bromwich inversion to find an expression for the 0 0.5 1 t / td 1.5 2
variable defined in (24)that depends on the parameter r, which
is the ratio of diffusivities and storage moduli inside and outside B) pore pressure evolution
,
the fault zone: long timescale
2 2
0. 1 ∆p ′(t) − ∆ p h + h N 2
k0 0 td =
hAN c0 N 2 ∆ p ′(0 + ) − ∆ p 8c ′ 8c N ′
@
r¼ . ¼ :
k N0 c N0 r =0
h 1
0.01
0.11
[48] Once the solution is found (equation(B2) in Appendix B), 0.8
1
we estimate what parameters control the timescale over which the c′ N 2 4
fault zone reaches pressure equilibrium with its surroundings. In r=
c N ′ ∞
Appendix B we show that the time at which the pore pressure at the 0.6
center of the fault has evolved approximately half way toward its
longer time limit of p is given by
0.4
h2 h2 N 2 h2
td ¼ 0 þ ¼ 0 ð1 þ rÞ: ð25Þ 0.2
8c 8c N 0 8c
[49] This time constant is the sum of a characteristic drainage 0
time for the fault core plus a characteristic time for the crust. Two 0 5 10 t / td 15 20
limit cases exist: whenr=0, the crust is much more permeable
than the fault zone, and more interesting ly, when r!1, the fault Figure 7. Temporal evolution of normalized pore pressure
zone is much more permeable than the surrounding crust. In the alteration at a (a) short and (b) long timescales for different
former case the time constant is simply td=h2/8c0, while the latter values of the parameter r(see text for its definition). The time
condition yields variable is normalized using the characteristic time for pore
pressure equilibrium, while the pore pressure alteration is
h2 N 2 normalized at its initial value at the instant of application of
td ¼ : the induced stress.
8c N 0
[50] In Figure 7 we show the temporal evolution of the
normalized pore pressure alteration (defined in (24)) for short representing a slow temporal decay as t1/2. This slow asymptotic
and long timescales and for different values of the parameter r. decay rate is independent of the permeability within the fault
Figure 7 shows that the solution for r4 is almost identical to zone, although the time at which that asymptotic expression
that obtained for r = 1. The pore pressure alteration at the fault becomes accurate does depend on fault permeability, since it
center is halfway toward its longer time limit after a time which enters in the ratio r. Infact, that asymptotic form, which is
varies from 0.75 td when r = 0, to 1.20 td when r = 1. This instructively written as
property is what motivated our definition of td. We may also note
that for all cases except r = 0, there is a slow evolution at long N p0 ð0þ Þ p
p0 ðt Þ p ! pffiffiffiffiffiffiffi h;
times, and thus the process is not readily characterized solely by 2 pct N0
td. Indeed, the asymptotic evaluation of the solution (B2) for large
t shows that is valid in the generalized form
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Z
p0 ðt Þ p 2rtd h N N þh=2
p0 ðz; 0þ Þ p
! ¼ pffiffiffiffiffiffiffi ; p0 ðt Þ p ! pffiffiffiffiffiffiffi dz ð26Þ
p0 ð0þ Þ p pð1 þ rÞt 2 pct N 0 2 pct h=2 N 0 ð zÞ
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X - 13
when we do not assume that the poroelastic material properties are Table 1. Characteristic Times for Local Pressure Equilibrium
uniform within the fault zone. Rather, in (26) we assume that all
Fault Thickness h, m Permeability
material properties vary with position z within the fault zone, such
asN 0 = N 0(z)but are uniform outside it (atjzj>h/2). Here k = 1018 m2 k = 1021 m2
4
p = Bskk/3 as before in the adjoining crust, and (26)applies 0.001 2.5 10 s 0.25 s
on a sufficiently long timescale that local pore pressure equilibrium 0.01 2.5 102 s 25.0 s
is achieved within the fault zone, with p0(t) denoting that value. 0.1 2.5 s 42 min
Also, p0(z, 0+) is the pore pressure change that would have been 1. 4.2 min 2.9 days
induced at position z (which coincides with the axis 3 in Figure 1) 10 6.9 hours 9.7 months
100 29 days 80 years
within the fault zone in undrained response to the stress change. It is 1000 8.0 years 80 centuries
given by our earlier expression for p0 as a linear combination of
skk/3 and s33 but now written as
Ku0 ð zÞ G0 ðzÞ Mu skk G G0 ð zÞ
p0 ðz; 0þ Þ ¼ B0 ð zÞ þ s33 and those of its surroundings. We have shown that if the rigidity
Mu0 ð zÞ G Ku 3 G
inside the fault zone is much smaller than that in the surrounding
crust, which should correspond to a large reduction of S wave
This allows, for example, for a fault core and bordering zone with velocity within the fault zone, the pore pressure is primarily
damage degrading gradually toward that appropriate for the controlled by the fault-normal stress changes and the definition
surrounding crust. of the effective friction coefficient given in literature is tenable.
[51] It is interesting to provide a tentative evaluation of the However, we emphasize that if the fault zone is an isotropic
characteristic time values for local pressure equilibrium. It poroelastic medium permeated by fluids, the fault-normal stress
emerges from (25) that the characteristic time will be dominated changes are the most important factor controlling the pore pressure
by the smallest value of diffusivity, c or c0, which effectively changes only when the reduction in S wave velocity is >50%. A
means the smallest permeability, since it is reasonable to assume limiting case, for which only the mean stress (i.e., the first
that the storage modulus and the fluid viscosity do not signifi- invariant) controls the pore pressure, is found for faults whose
cantly change between the fault zone and the surrounding crust. rigidity is equal to that of the surrounding crust. However, such a
In Table 1 we list the values of td = h2/8cmin, understanding for condition appears inconsistent with observations of fault zone
cmin the smaller diffusivity value. We assume a fluid viscosity of structure resulting from seismic tomography and fault zone trap-
2 104 Pa s and a storage modulus of 100 GPa. We have ped-wave studies. These studies consistently show that body wave
calculated the characteristic times for two values of permeability: velocities within the fault zone are different from those in the
1018 m2 and 1021 m2, respectively, and for fault thickness in surrounding crust. In particular, fault zone trapped-wave studies
the interval 1 mm h 1 km. As expected, the resulting values indicate that shear wave velocities in fault zones are as much as
of the characteristic time for local pressure equilibrium change 50% smaller than in their surroundings. If the fault zone materials
from seconds to years. This suggests that it is really difficult to do not behave as in these extreme conditions, both the mean stress
exclude a priori the contribution of time-dependent pore pressure and the fault-normal stress changes contribute together to the pore
equilibrium in the analysis of stress redistribution. Moreover, a pressure changes for undrained deformation during such a short
complex fault network will inevitably have segments at different timescale.
stages of their relaxation from undrained conditions to local [54] Calculations of Coulomb stress changes caused by shear
pressure equilibrium with the nearby materials. dislocations in an elastic isotropic half-space show that the
choice of the pore pressure model influences the results
significantly. In particular, we show that the use of a constant
8. Discussion and Concluding Remarks effective friction model (equations (3) and (4)) as opposed to
an isotropic and homogeneous pore pressure model (equations
[52] Postseismic stress redistribution is a time-dependent (1) and (2)) implies very different Coulomb stress changes [see
process, and at short or intermediate timescales (from minutes also Beeler et al., 2000]. These stress changes also depends on
to few years after a seismic event), fluid flow can be one of the the assumed values of friction and Skempton parameters.
most important factors in contributing to this temporal variation [55] We also briefly investigated the effect of anisotropy in the
of the stress perturbation [Nur and Booker, 1972; Hudnut et al., cracked region forming the fault core. In this case, the Skempton
1989; Noir et al., 1997]. In order to properly include a pore parameter becomes a tensor. However, it is plausible to assume that
pressure model in Coulomb analysis through equation (1) it is the strain component normal to the fault is much larger than those
necessary to choose the timescale during which the stress in the fault-parallel directions. In that case, the pore pressure
changes are modeled as well as to make a few assumptions should be solely dependent on the fault-normal stress, and the
on the material properties of the medium. In this study we have definition of the effective friction coefficient should be correct.
investigated two different timescales. We have first focused our Thus we can conclude that at very short postseismic time periods in
attention on the short-term postseismic period, in which both the which the fault zone obeys undrained conditions, the constant
fault zone and the adjoining lithosphere respond under undrained effective friction model could be an acceptable approximation only
conditions. Thus we neglect the alteration of pore pressure under quite extreme conditions, such as if the fault zone rigidity is
caused by fluid flow. In this first configuration we discuss the <50% that of its surroundings or if the fault zone is strongly
pore pressure changes both in an isotropic poroelastic medium anisotropic. This latter configuration would be approached if
and in an anisotropic fault zone. porosity is dominated by an oriented distribution of cracks or
[53] That first condition allows a comparison with most of the flattened pores aligned with their long directions subparallel to the
Coulomb stress studies. We have derived an analytical expression fault plane.
that relates pore pressure changes for undrained deformation within [56] We have also discussed an intermediate timescale, which
the fault zone to mean and fault-normal stress changes for an will exist only for a sufficiently permeable fault that is locally
isotropic poroelastic medium. Both these terms contribute to the drained and reaches pressure equilibrium with its surroundings. In
variations of pore pressure caused by the stress redistribution this time period the adjoining lithosphere is still responding as if it
process. Their relative weight in the derived equation depends on were effectively undrained. In this case, we assume that the fault
the contrast between the elastic parameters inside the fault zone zone is moderately thin and has some permeability. During this
ESE X - 14 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
relatively short timescale, p0 evolves toward p, the former where p0 is an unessential reference pressure, we may use the
being time-dependent due to the pore fluid fluxes generated by property that d m̂ = dp/r to obtain
gradients in the coseismically induced pore pressure field. This
variation of P in the adjoining rock will be slower than the local
m pm m
drained response of the fault. Therefore, on such a short but not pd ¼d dp ðA3Þ
r r r
extremely short timescale, during which the fault acts as drained,
but its surroundings do not, we get p0 p, and thus p0 is pm pm
¼d md ðm
^Þ ¼ d m
^m þ m
^dm:
proportional to the mean stress changes. The characteristic time of r r
this local pressure equilibrium depends on the permeability and the
fault thickness. Thus (A1) can be transformed to another perfect differential
[57] The transition from short-term undrained to drained which is well known in the thermodynamics of mixtures,
response in the adjoining lithosphere occurs in general on a yet namely,
longer timescale. This should always occur as time increases,
except when the fault zone is hydrologically isolated from its
surroundings. In these circumstances, the stress changes sij m
sij deij þ m
^dm ¼ d U þ ^
mm p : ðA4Þ
approach the values that would be calculated from elastic disloca- r
tion theory using drained, rather than undrained, elastic moduli.
The drained response is elastically less stiff than the undrained [61] While it is unessential for what follows, the expression
response: K < Ku, v < vu, G = Gu. In general, we might expect that of (A2) for m̂ may be seen to be consistent with interpreting
this transition from undrained to drained response modestly m̂dm as the total reversible work of extracting an element of
reduces the stress changes. For a mode II shear crack, in a plane mass dm from a reservoir of fluid at a reference pressure and
strain condition, the reduction scales as (1 vu)/(1 v). For fluid density ( p0, r0 = r(p0)), and inserting it (say, through a
Westerly granite this value is close to 0.9; thus the stress reduction porous screen) into a porous medium at a place where the
at longer time periods seems almost negligible. However, a pore pressure and fluid density are (p, r). We assume that
complete recognition of this behavior during such longer timescale temperature is the same in the reservoir as in the place of
is beyond the aims of the present study. insertion and calculate the work in three steps, as follows: (1)
[58] The time dependence is included in the Coulomb failure work of withdrawal from reservoir, p0dm/r0 (note that dm/r0
function not only through the pore pressure changes p (equation is the volume withdrawn from the reservoir),R (2) work of
(1)). In fact, poroelastic theory shows that there is a time changing density from r0 to r, which is dm r0rpd(1/r). (3)
dependence of all the stress components sij (we have considered Work of inserting the fluid at the place where pressure is p,
here only timescales for which those sij are effectively constant which is pdm/r (dm/r is the volume inserted). The sum is
outside the fault zone). Nur and Booker [1972] pointed out that m̂dm, so that
the time dependence of pore pressure changes can interact with
seismicity explaining aftershocks, and Rice [1980] evaluated the Zr Zp
resulting time-dependent postseismic shear stress history on a dm dm
fault surface. Further observations are needed to image the inner mdm ¼ po
^ dm pd ð1=rÞþp ¼ dm ð1=rÞdp; ðA5Þ
ro r
structure of fault zones and therefore to verify the most appro- ro po
priate pore pressure model. In absence of constraining evidence
we cannot exclude any model for including pore pressure in which is consistent with the expression for m̂ in (A2).
Coulomb failure. However, at least in a homogeneous and [62] By a final rearrangement, we obtain
isotropic poroelastic medium, the influence on pore pressure of
the mean stress is well established. eij dsij þ m
^dm ¼ dV ðA6Þ
as a perfect differential, where V = U + m̂m p(m/r) sijeij.
Appendix A. Induced Pore Pressure for Regarding V as a function of the set of stresses [s] and fluid mass
Undrained Stressing of an Anisotropic m, V = V([s], m), it therefore follows that
Medium and Interpretation
of a Skempton Tensor @V ð½s
; mÞ @V ð½s
; mÞ
^¼
m ; eij ¼ : ðA7Þ
[59] An increment of work (per unit volume of the reference @m @sij
state) done on a poroelastic material, precisely on its solid phase, is
given by sijdeij + pd(m/r)[e.g., Rice and Cleary, 1976], where m/r Recognizing that @ 2V([s], m)/@mdsij must be independent of the
is the fluid volume fraction defined as in the text. We recall that m order of differentiation, the Maxwell reciprocal relation
is the mass of fluid per unit reference state volume of porous
material and r .is the density of pure fluid at pressure p, and we
mð½s
; mÞ
@^ @eij ð½s
; mÞ
assume r = r(p). This increment of work must be a perfect ¼ ðA8Þ
differential of a function of state (i.e., of the strain energy U of @sij @m
the solid phase, or of its Helmholtz free energy, at the constant
temperature conditions considered), and so must be valid. Its left side can be rewritten as
sij deij þ pd ðm=rÞ ¼ dU : ðA1Þ mð½s
; mÞ
@^ mð pÞ @pð½s
; mÞ 1 @pð½s
; mÞ
@^
¼ ¼ ; ðA9Þ
@sij @p @sij r @sij
[60] Introducing the ‘‘chemical potential’’ m̂ = m̂(p) by
and so the Maxwell relation is equivalent to
Zp
1
^¼m
m ^ð pÞ ¼ d^p; ðA2Þ @pð½s
; mÞ @eij ð½s
; mÞ
rð ^pÞ ¼ rð pÞ : ðA10Þ
p0 @sij @m
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X - 15
[63] We recognize that @p([s], m)/@sij as being a generalization since N = ch/k. Thus, from equation (17) of Rice and Cleary
of the Skempton coefficient B, valid for the anisotropic case as [1976], but with their expressions in terms of Poisson ratios
well, and make the definition rewritten in terms of moduli used earlier here, we have N =
B2Ku2M/(Ku K)Mu.
Bij @pð½s
; mÞ [66] Measuring P relative to its value at time t = 0 (just before
¼ ðA11Þ the earthquake), so that p = p0 in the fault zone and p outside in
3 @sij the crust at t = 0+, and letting
to define a Skempton tensor, with property that p = Bsij/ Z1
3 under undrained conditions. (Of course, in the isotropic case,
pðz; sÞ ¼
^ pðz; sÞest dt
B ij = Bdij .) The Maxwell relation then gives us an
interpretation of, and alternative way of understanding, the 0
Skempton tensor as
be the Laplace transform, the solution of the above equation set is
@eij ð½s
; mÞ pffiffiffiffiffiffiffiffi
Bij ¼ 3r : ðA12Þ dp0 ð0þ Þ ½p0 ð0þ Þ p
cosh z s=c0
@m ^pðz; sÞ ¼ pffiffiffiffiffiffiffiffi pffiffi pffiffiffiffiffiffiffiffi
zjzj < h=2
s s cosh ðh=2Þ s=c0 þ r sinh ðh=2Þ s=c0
Appendix B. Short-Term Pore Pressure
pffiffi pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
Equilibrium Between Fault Zone ^pðz; sÞ ¼
p ½p0 ð0þ Þ p
r sinh ðh=2Þ s=c0 exp ðjzj ðh=2ÞÞ s=c0
þ pffiffiffiffiffiffiffiffi pffiffi pffiffiffiffiffiffiffiffi
jzj > h=2;
and Surrounding Crust s 0
s cosh ðh=2Þ s=c þ r sinh ðh=2Þ s=c 0
[64] The fault is modeled as a zone of thickness h (see Figure 1),
which has uniform permeability k0, fluid viscosity h0, and storage where now p0(0+) is the same as p0 of (13), and the parameter:
modulus N0 (defined below). The surrounding crust is modeled as a 0
pair of semi-infinite domains with corresponding parameters k, h, k
h0 N c0 N 2
N. Our analysis of short-time undrained response shows that at r¼ ¼ :
time t = 0+ the fault has the pore pressure change p0 (which we k 0
c N0
N
have calculated in (13) as B0 times a linear combination of s33/ h
3 and s33), whereas the surrounding crust has the change p =
Bskk/3. [67] It is simplest to invert the transform solution for the pore
[65] The poroelastic equations then allow us to model the pressure at the center of the fault zone, and we use the notation
evolution of the pore pressure in the fault toward its longer time
limit p = Bskk/3. Recognizing that only e33 and no other
strain varies with time, and noting that s33 is uniform, the same p0 ðtÞ pð0; t Þ;
inside and outside the fault, the problem is recognized as one of
one-dimensional consolidation. The governing equations within so that
the fault and the crustal domains, respectively, incorporating
Darcy’s law and conservation of mass of the diffusing fluid, for t ^p0 ðsÞ ^
pð0; sÞ:
> 0 take the forms of
Then
k 0 dpðz; tÞ 1 dpðz; tÞ
¼ 0 ; h=2 < z < h=2 !
dz h0 dz N dt p0 ðsÞ p=s 1
^ 1
1 pffiffiffiffiffiffiffiffi pffiffi pffiffiffiffiffiffiffiffi :
d dpðz; tÞ 1 dpðz; tÞ p0 ð0þ Þ p s cosh ðh=2Þ s=c0 þ r sinh ðh=2Þ s=c0
kh ¼ ; z > h=2; z < h=2;
dz dz N dt
The Bromwich inversion integral is then
where z is the spatial coordinate in the 3 direction, perpendicular to
the fault. These are homogeneous diffusion equations for p. The 0þ
general consolidation equations instead involve a homogeneous Zþi1
p0 ðtÞ p 1 ^p0 ðsÞ p=s st
diffusion equation for m [Rice and Cleary, 1976], not p, but reduce ¼ e ds: ðB1Þ
p0 ð0þ Þ p 2pi p0 ð0þ Þ p
to such an equation for P in the case of one-dimensional straining, 0þ i1
like here. The diffusivities c0 = k0N0/h0 and c = kN/h. Their solution
will be even in z and must satisfy, for t > 0, Since the integrand vanishes rapidly enough as jsj ! 1 and has no
poles (at least when r > 0) but has a branch cut along the negative
þ real S axis, the inversion path can be distorted to run from 1 to 0
h h along the lower side of the cut and from 0 to 1 along the upper
p ;t ¼ p ;t
2 2 side. We make the substitution s = 4c0x2/h2 in the inversion
k 0 dpðz; t Þ k dpðz; t Þ integral, where x is real and nonnegative along the distorted
¼
h0 dz h
z¼h dz
z¼hþ inversion path, and note that
2 2
est ¼ exp 4c0 t=h2 ¼ exp ð1 þ rÞx2 t=2td ;
The storage modulus N (inverse of the storage coefficient) is
defined such that dm = r0dp/N is the increase in fluid mass content where
when the pressure varies under conditions of one-dimensional
strain, with s33 held constant. Expressions for it can be extracted
h2 h2 N 2
from those for c in the sources mentioned on poroelasticity [Biot, td ¼ þ :
1941, 1956; Rice and Cleary, 1976; Kuempel, 1991; Wang, 2000], 8c0 8c N 0
ESE X - 16 COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS
We find from numerical evaluations that td gives the time at which Eberhart-Phillips, D., and A. J. Michael, Three-dimensional velocity struc-
the pore pressure at the center of the fault has evolved ture, seismicity, and fault structure in the Parkfield region, central Cali-
fornia, J. Geophys. Res., 98, 15,737 – 15,758, 1993.
approximately halfway toward its longer time limit of p = Eberhart-Phillips, D., and M. Reyners, Plate interface properties in the
Bskk/3. Hence northeast Hikurangi subduction zone, New Zealand, from converted seis-
mic waves, Geophys. Res. Lett., 26, 2565 – 2568, 1999.
pffiffi Z1 ! Green, D. H., and H. F. Wang, Fluid pressure response to undrained com-
p0 ðt Þ p 2 r sinð xÞexp½ð1 þ rÞx2 t=2td pression in saturated sedimentary rock, Geophysics, 51, 948 – 956, 1986.
¼ dx: ðB2Þ Harris, R. A., Introduction to special section: Stress triggers, stress sha-
p0 ð0þ Þ p p x cos2 ð xÞ þ r sin2 ð xÞ
0 dows, and implications for seismic hazard, J. Geophys. Res., 103,
24,347 – 24,358, 1998.
Harris, R. A., and S. M. Day, Dynamics of fault interaction: parallel strike-
The integral converges rapidly for t > 0 and can be evaluated by slip faults, J. Geophys. Res., 98, 4461 – 4472, 1993.
standard numerical integration schemes, although the number of Harris, R. A., and R. W. Simpson, Changes in static stress on southern
numerical subdivisions becomes very large when r is either very California faults after the 1992 Landers earthquake, Nature, 360, 251 –
large or small compared to unity. Fortunately, the limit cases can be 254, 1992.
treated separately. We find that when r = 0 (crust very much more Hart, D. J., Laboratory measurements of a complete set of poroelastic
moduli for Berea sandstone and Indiana limestone, M. S. thesis, 44
permeable than the fault zone), the branch cut changes into a row pp., Univ. of Wis., Madison, 1994.
of simple poles, giving Hill, D. P., et al., Seismicity remotely triggered by the magnitude 7.3
Landers, California, earthquake, Science, 260, 1617 – 1623, 1993.
Hudnut, K. W., L. Seeber, and J. Pacheco, Cross-fault triggering in the
p0 ðt Þ p 4X1
ð1Þn 2
2p t November 1987 Superstition Hills earthquake sequence, southern Cali-
¼ exp ð 2n þ 1Þ r ¼ 0;
p0 ð0þ Þ p p n¼0 2n þ 1 8td fornia, Geophys. Res. Lett., 16, 199 – 202, 1989.
Johnson, P. A., and T. V. McEvilly, Parkfield seismicity: Fluid driven?, J.
Geophys. Res., 100, 12,937 – 12,950, 1995.
in which case td = h2/8c0. That is just the classical Terzaghi one- King, G. C. P., and M. Cocco, Fault interaction by elastic stress changes:
dimensional consolidation solution for a layer which is freely New clues from earthquake sequences, Adv. Geophys., 44, 1 – 38, 2000.
drained at both sides. King, G. C. P., and R. Muir-Wood, The impact of earthquakes on fluids in
the crust, Ann. Geofis., XXXVII(6), 1453 – 1460, 1994.
p[68] Also, either direct treatment or use of the substitution y = King, G. C. P., R. S. Stein, and J. Lin, Static stress changes and the
x r and taking the limit r ! 1 (fault zone very much more triggering of earthquakes, Bull. Seismol. Soc. Am., 84, 935 – 953, 1994.
permeable than crust) gives Kuempel, H. J., Poroelasticity: Parameters reviewed, Geophys. J. Int., 105,
783 – 799, 1991.
Z1 Leary, P. C., Y. G. Li, and K. Aki, Observation and modelling of fault-zone
p0 ðt Þ p 2 expðy2 t=2td Þ fracture seismic anisotropy, I, P, SV and SH travel times, Geophys. J.
¼ dy Astron. Soc., 91, 461 – 484, 1987.
p0 ð0þ Þ p p 1 þ y2
0 r ¼ 1; Lees, J. M., Tomographic P-wave velocity images of the Loma Prieta
earthquake asperity, Geophys. Res. Lett., 17, 1433 – 1436, 1990.
Lees, J. M., and C. E. Nicholson, Three-dimensional tomography of the 1992
in which case, southern California earthquake sequence, Geology, 21, 387 – 390, 1993.
Li, Y. G., and P. C. Leary, Fault zone trapped seismic waves, Bull. Seismol.
h2 N 2 Soc. Am., 80, 1245 – 1271, 1990.
td ¼ : Li, Y. G., P. C. Leary, K. Aki, and P. Malin, Seismic trapped modes in the
8c N 0 Oroville and San Andreas fault zones, Science, 249, 763 – 766, 1990.
Li, Y. G., J. Vidale, K. Aki, J. Marone, and W. K. Lee, Fine structure of the
The above expressions have been plotted in Figure 7 for various Landers fault zone: Segmentation and the rupture process, Science, 265,
values of r on short and long timescales. 367 – 380, 1994.
Mooney, W. D., and A. Ginzburg, Seismic measurements of the internal
properties of fault zones, Pure Appl. Geophys., 124, 141 – 157, 1986.
Noir, J., E. Jacques, S. Békri, P. M. Adler, P. Tapponnier, and G. C. P. King,
[69] Acknowledgments. J.R.R. wishes to acknowledge the support of Fluid flow triggered migration of events in the 1989 Dobi earthquake
the NSF Geophysics Program and the USGS Earthquake Hazards Reduc-
sequence of central Afar, Geophys. Res. Lett., 24, 2335 – 2338, 1997.
tion Program and also of a Blaise Pascal International Research Chair from
Nostro, C., M. Cocco, and M. E. Belardinelli, Static stress changes in
the Foundation of École Normale Supérieure, Paris. M.C. wishes to
extensional regimes: An application to southern Apennines (Italy), Bull.
acknowledge the Institute de Physique du Globe de Paris for the hospitality
Seismol. Soc. Am., 87, 234 – 248, 1997.
during his stay in France. This study was partially supported by the
Nur, A., and J. R. Booker, Aftershocks caused by pore fluid flow?, Science,
European Commission contract ENV-4970528, Project Faust. We thank
175, 885 – 887, 1972.
T. Yamashita and an anonymous referee for their comments and criticism
O’Connell, R. G., and B. Budiansky, Seismic velocities in dry and saturated
that helped in improving the manuscript.
cracked solids, J. Geophys. Res., 79, 5412 – 5426, 1974.
Okada, Y. , Surface deformation due to shear and tensile faults in a half-
space, Bull. Seismol. Soc. Am., 75, 1135 – 1154, 1985.
References Okada, Y., Internal deformation due to shear and tensile faults in a half-
Beeler, N. M., R. W. Simpson, D. A. Lockner, and S. H. Hickman, Pore space, Bull. Seismol. Soc. Am., 82, 1018 – 1040, 1992.
fluid pressure, apparent friction and Coulomb failure, J. Geophys. Res., Parsons, T., R. S. Stein, R. W. Simpson, and P. Reasenberg, Stress sensi-
105, 25,533 – 25,554, 2000. tivity of fault seismicity: A comparison between limited-offset oblique
Belardinelli, M. E., M. Cocco, O. Coutant, and F. Cotton, Redistribution of and major strike-slip faults, J. Geophys. Res., 104, 20,183 – 20,202, 1999.
dynamic stress during coseismic ruptures: Evidence for fault interaction Perfettini, H., R. Stein, R. Simpson, and M. Cocco, Stress transfer by the
and earthquake triggering, J. Geophys. Res., 104, 14,925 – 14,945, 1999. M = 5.3, 5.4 Lake Elsnam earthquakes to the Loma Prieta fault:
Biot, M. A., General theory of three-dimensional consolidation, Appl. Unclamping at the site of the peak 1989 slip, J. Geophys. Res., 104,
Phys., 12, 155 – 164, 1941. 20,169 – 20,182, 1999.
Biot, M. A., General solutions of the equations of elasticity and consolida- Rice, J. R., The mechanics of earthquake rupture, in Physics of the Earth
tion for a porous material, J. Appl. Mech., 78, 91 – 98, 1956. Interior, edited by A. M. Dziewonski and E. Boschi, Proc. Int. Sch. Phys.
Byerlee, J. D., The change in orientation of subsidiary shears near faults Enrico Fermi, 78, 555 – 649, 1980.
containing pore fluid under high pressure, Tectonophysics, 211, 295 – Rice, J. R., Fault stress states, pore pressure distributions and the weakness
303, 1992. of the San Andreas fault, in Fault Mechanics and Transport Properties of
Cocco, M., C. Nostro, and G. Ekstrom, Static stress changes and fault Rock, edited by B. Evans and T.-F. Wong, pp. 475 – 503, Academic, San
interaction during the 1997 Umbria-Marche earthquake sequence, J. Seis- Diego, Calif., 1992.
mol., 4, 501 – 516, 2000. Rice, J. R., and M. P. Cleary, Some basic stress diffusion solutions for fluid-
Cotton, F., and O. Coutant, Dynamic stress variations due to shear faulting saturated elastic porous media with compressible constituents, Rev. Geo-
in a plane-layered medium, Geophys. J. Int., 128, 676 – 688, 1997. phys., 14, 227 – 241, 1976.
COCCO AND RICE: POROELASTICITY EFFECTS IN COULOMB STRESS ANALYSIS ESE X - 17
Roeloffs, E., Poroelastic techniques in the study of earthquake-related hy- Thurber, C. S., W. Roecker, W. Ellsworth, Y. Chen, W. Lutter, and R.
drologic phenomena, Adv. Geophys., 37, 135 – 195, 1996. Sessions, Two-dimensional seismic image of the San Andreas fault in
Roeloffs, E., Persistent water level changes in a well near Parkfield, Cali- the northern Gabilan Range, central California: Evidence for fluids in the
fornia, due to local and distant earthquakes, J. Geophys Res., 103, 869 – fault zone, Geophys. Res. Lett., 24, 1591 – 1594, 1997.
889, 1998. Wang, H. F., Theory of Linear Poroelasticity, Princeton Univ. Press,
Roeloffs, E., and E. G. Quilty, Water level and strain changes preceding and Princeton, N. J., 2000.
following the August 4, 1985, Kettleman Hills, California, earthquake, Zhao, D., and H. Kanamori, The 1992 Landers earthquake sequence: Earth-
Pure Appl. Geophys., 149, 21 – 60, 1997. quake occurrence and structural heterogeneities, Geophys. Res. Lett., 20,
Roeloffs, E., and J. W. Rudnicki, Coupled deformation diffusion effects on 1083 – 1086, 1993.
water level changes due to propagating creep events, Pure Appl. Geo- Zhao, D., and H. Kanamori, The 1994 Northridge earthquake: 3-D crustal
phys., 122, 560 – 582, 1985. structure in the rupture zone and its relation to the aftershock locations
Scholz, C. H., The Mechanics of Earthquakes and Faulting, 439 pp., Cam- and mechanisms, Geophys. Res. Lett., 22, 763 – 766, 1995.
bridge Univ. Press, New York, 1990. Zhao, D., and T. Mizuno, Crack density and saturation rate in the
Segall, P., and J. R. Rice, Dilatancy, compaction, and slip instability of a 1995 Kobe earthquake region, Geophys. Res. Lett., 26, 3213 – 3216,
fluid infiltrated fault, J. Geophys. Res., 100, 22,155 – 22,171, 1995. 1999.
Sibson, R. H., Crustal stress, faulting and fluid flow, in Geofluids: Origin, Zhao, D., and H. Negishi, The 1995 Kobe earthquake: Seismic image of the
Migration and Evolution of Fluids in Sedimentary Basins, edited by J. source zone and its implications for the rupture nucleation, J. Geophys.
Parnell, Geol. Soc. Spec. Publ., 78, 69 – 84, 1994. Res., 103, 9967 – 9986, 1998.
Simpson, R. W., and P. A. Reasenberg, Earthquake-induced static stress Zhao, D., H. Kanamori, and H. Negishi, Tomography of the source area of
changes on central California faults, in The Loma Prieta, California the 1995 Kobe earthquake: Evidence for fluids at the hypocenter?,
Earthquake of October 17, 1989—Tectonic Processes and Models, edited Science, 274, 1891 – 1894, 1996.
by R. W. Simpson, U.S. Geol. Surv. Prof. Pap., 1550-F, F55 – F89, 1994.
Skempton, A. W., The pore pressure coefficients A and B, Geotechnique, 4,
143 – 147, 1954.
Stein, R. S., The role of stress transfer in earthquake occurrence, Nature, M. Cocco, Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna
402, 605 – 609, 1999. Murata 605, I-00143 Rome, Italy. (cocco@ingv.it)
Stein, R. S., G. C. P. King, and J. Lin, Change in failure stress on the J. R. Rice, Engineering Sciences and Geophysics, Harvard University, 29
southern San Andreas fault system caused by the 1992 magnitude = Oxford St, 224 Pierce Hall, Cambridge, MA 02138, USA. (rice@
7.4 Landers earthquake, Science, 258, 1328 – 1332, 1992. esag.harvard.edu)
ERRATA CORRIGE (this compilation: 4 October 2002)
1. James R. Rice's correct affiliation (first page, just below title) is:
Department of Earth and Planetary Sciences and Division of Engineering and Applied
Sciences, Harvard University, Cambridge, Massachusetts.
2. There is a mistake in the equation included in Figures 4 and 5 for the isotropic
poroelastic model. The correct equation is: ∆CFF = ∆τ + µ(∆σ n − Β∆σ kk / 3) . The
figures were computed with the proper sign and are correct; only that equation is
misprinted.
3. Paragraph [38], Equation (21) should be: Bij = 3∂p([σ ], m) / ∂σ ij (correcting the
placement of the closing square bracket and eliminating the unnecessary curly brackets).
4. Paragraph [40], Equation (22) should be: Bij = 3ρ∂εij ([σ ], m) / ∂m (again, correcting the
placement of the closing square bracket and eliminating the unnecessary curly brackets).
5. End of paragraph [44]: The correct relation is: ∆p′ = −Β∆σ kk / 3 (instead of
∆p′ ≈ −∆σ kk / 3).
6. Appendix A, Equation (A3): The expression following the last equal sign should be:
pm
d − µˆ m + µˆ dm
ρ
7. Appendix B, paragraph [65]: The upper case P which appears between the two sets of
equations in that paragraph should instead be lower case p.
8. Appendix B, paragraph [65]: Last member of second set of equations has the lower
case deltas within both parentheses. All four such deltas should be the partial derivative
sign ( δ should be replaced by ∂ ).
9. Appendix B, paragraph [66]: The first term on the right of the second equation has a
lower case delta in the numerator which should be an upper case delta ( δ should be
replaced by ∆ ).
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B2, 2069, doi:10.1029/2002JB002319, 2003
Correction to ‘‘Pore pressure and poroelasticity effects in Coulomb
stress analysis of earthquake interactions’’
by Massimo Cocco and James R. Rice
Received 22 November 2002; published 4 February 2003.
INDEX TERMS: 9900 Corrections; KEYWORDS: fault interaction, fluid flow, poroelasticity, effective friction,
crustal anisotropy
Citation: Cocco, M., and J. R. Rice, Correction to ‘‘Pore pressure and poroelasticity effects in Coulomb stress analysis of earthquake
interactions’’ by Massimo Cocco and James R. Rice, J. Geophys. Res., 108(B2), 2069, doi:10.1029/2002JB002319, 2003.
[1] In the paper ‘‘Pore pressure and poroelasticity effects 4. Paragraph [40], Equation (22) should be Bij = 3r@eij([s],
in Coulomb stress analysis of earthquake interactions’’ by m)/@m (again, correcting the placement of the closing bracket
Massimo Cocco and James R. Rice (Journal of Geophysical and eliminating the unnecessary curly brackets).
Research, 107(B2), 2030, doi:10.1029/2000JB000138, 5. End of paragraph [44]: The correct relation is p0 =
2002), there are several corrections as follows: Bskk/3 (instead of p skk/3).
1. James R. Rice’s correct affiliation (first page, just 6. Appendix A, equation (A3): The expression
following
below title) should be Department of Earth and Planetary the last equals sign should be d pm
r m
^ m þ m
^ dm.
Sciences and Division of Engineering and Applied 7. Appendix B, paragraph [65]: The capital P that appears
Sciences, Harvard University, Cambridge, Massachusetts. between the two sets of equations in that paragraph should
2. There is a mistake in the equation included in Figures instead be lowercase p.
4 and 5 for the isotropic poroelastic model. The correct 8. Appendix B, paragraph [65]: Last member of second set
equation is CFF = t + m(sn Bskk/3). The figures of equations has the lower case deltas within both
were computed with the proper sign and are correct, only parentheses. All four such deltas should be the partial
the text is wrong. derivative sign (d should be replaced by @).
3. Paragraph [38], Equation (21) should be Bij = 3@p([s], 9. Appendix B, paragraph [66]: The first term on the right
m)/@sij (correcting the placement of the closing bracket and of the second equation has a lowercase delta in the numerator
eliminating the unnecessary curly brackets). which should be an capital delta (d should be replaced by ).
Copyright 2003 by the American Geophysical Union.
0148-0227/03/2002JB002319$09.00
ESE 2-1