Sunteți pe pagina 1din 13

Geochemistry

Geophysics
Geosystems

Article
Volume 6, Number 11
15 November 2005
Q11008, doi:10.1029/2005GC000991

AN ELECTRONIC JOURNAL OF THE EARTH SCIENCES


Published by AGU and the Geochemical Society

ISSN: 1525-2027

Physical basis of evolution laws for rate and state friction


Norman H. Sleep
Department of Geophysics, Stanford University, Mitchell Building, Room 360, 397 Panama Mall, Stanford, California
94305, USA (norm@geo.stanford.edu)

[1] The well-known formalism of rate and state dependent friction represents the transition between
starting friction and sliding friction. It expresses the instantaneous coefficient of friction in macroscopic
quantities as m = m0 + a ln (V/V0) + b ln (y/ynorm), where m0 is the first-order coefficient of friction, a and b
are small (0.01) dimensionless constants, V is sliding velocity, V0 is a reference sliding velocity, and the
inverse of the state variable 1/y represents damage. The normalizing state variable ynorm  (P/P0)a/b
represents the effect of the normal traction P on stress concentrations, where P0 is a constant with
dimensions of pressure and a  b is another dimensionless constant. Evolution equations represent the
combined effect of damage from sliding and healing on the state variable. The Ruina (1983) evolution
equation implies that the state variable does not change (no healing) during holds when sliding is stopped.
It arises from exponential creep within gouge when the concentrated stress at asperities scales with the
macroscopic quantities. The parameter b  a being positive is a necessary condition for a spring-slider
system becoming unstable. Microscopically, this parameter represents the tendency of asperities
accommodating shear creep to persist longer than asperities of compaction creep at high sliding
velocities. In the Dieterich (1979) evolution law, healing occurs when the sample is at rest. It is a special
case where creep that produces shear and creep that produces compaction occur at different microscopic
locations at a subgrain scale. It also applies qualitatively for compaction at a shear traction well below that
needed for frictional sliding. Chemically, it may apply when shear sliding occurs within weak microscopic
regions of hydrated silica while compaction creep occurs within comparatively anhydrous grains.
Components: 7334 words, 4 figures.
Keywords: asperity; evolution laws; friction; rate and state.
Index Terms: 5104 Physical Properties of Rocks: Fracture and flow; 5120 Physical Properties of Rocks: Plasticity, diffusion,
and creep; 7209 Seismology: Earthquake dynamics (1242).
Received 5 April 2005; Revised 18 July 2005; Accepted 3 August 2005; Published 15 November 2005.
Sleep, N. H. (2005), Physical basis of evolution laws for rate and state friction, Geochem. Geophys. Geosyst., 6, Q11008,
doi:10.1029/2005GC000991.

1. Introduction
[2] The rate and state friction formalism describes
the transition between static friction and sliding
friction. It provides a good representation of laboratory data by including a term for damage associated with sliding and a term for healing when the
slider is slows down or is at rest. These terms result
in the attractive property that the formalism can
represent repeated earthquake cycles where the
fault strengthens during the interseismic period. It
Copyright 2005 by the American Geophysical Union

is well known that the second-order terms representing damage and healing determine mathematically whether a fault creeps peaceably or fails
catastrophically in an earthquake.
[3] I concentrate on healing in this paper. One
would like to know the physical basis of the laws
that purport to represent it before exporting laboratory results to the earthquake cycle at depth.
Troublingly, there are two different evolution laws
for keeping track of damage and healing on the
fault surface. The Dieterich [1979] evolution law
1 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

Figure 1. The paper considers stress involved in friction on several scales. Most laboratory experiments record the
sliding velocity and the stresses averaged over the contact surface. The stresses on a patch of the surface may differ
from these averages. The mesoscopic stress and strain rate within the gouge are useful for representing rate and state
equations as flow laws. The microscopic stress and strain rate within the small part of a grain, like an asperity of real
contact, are useful for representing deformation as thermally activated creep.

predicts that a fault heals when stopped so that its


friction when slipping is resumed is higher than
that before the hold. The Ruina [1983] evolution
law predicts no healing during a hold. An intriguing result is that the Ruina [1983] behavior occurs
at low relative humidity and the Dieterich [1979]
behavior at high relative humidity in simulated
gouge [Frye and Marone, 2002]. I address the
physical basis of rate and state friction with these
issues in mind.

2. Rate and State Equations


[ 4] The formalism for rate and state friction
evolved from laboratory experiments where one
measures (or controls) the macroscopic quantities
of normal traction P (in general effective pressure,
but I stick to drained conditions for simplicity),
shear traction t, and sliding velocity V (using the
notation of Sleep et al. [2000]). Scientists have
developed several simple semiempirical relationships between these measurable quantities and
porosity. They are approximations, to be sure, but
the simple forms allow revealing mathematical
manipulation. I begin with macroscopic properties
as would be measured over a few centimeter-squared
contact on a laboratory apparatus (Figure 1). I
introduce mesoscopic properties averaged over
enough grains that a continuum is meaningful in
section 2.1. I introduce microscopic quantities at

real contacts in section 2.2. They apply at scales


significantly greater than atomic dimensions where
a continuum is meaningful, that is, a few nanometers. The stress on a few millimeter wide patch
of the macroscopic contact arises in section 3.4.
[5] I begin with a unified theory of rate and state
friction compiled by Sleep [1997] and Sleep et al.
[2000] based on the Dieterich [1979] evolution
law. A state variable y represents the condition of
the surface or gouge; 1/y is a measure of damage.
The instantaneous shear traction from friction is in
terms of macroscopic variables
t Pm0 a lnV =V0 b lny=ynorm
;

where m0 is the coefficient of friction at reference


conditions, a and b are small dimensionless
constants, V0 is a reference velocity, and ynorm is
the normalizing value for the state variable, which I
discuss below. The state variable evolves with time
from Dieterichs [1979] law
@y V0 Pa=b yV

;

a=b
@t
Dc
Dc P

where t is time, Dc is the critical displacement to


significantly change the state properties of the
sliding surface, a is a dimensionless parameter
than represents the behavior of the surface after
a change in normal traction from Linker and
2 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

Dieterich [1992], and P0 is a reference normal


traction. The first right-hand term represents
healing and the second term damage from sliding.
It is well know that b > a is a necessary condition
for a spring-slider system (including a fault in the
elastic Earth) to become unstable.

2.1. Strain Rate Form of Equations


[6] The relationship of frictional sliding to other
flow laws becomes more apparent if one uses strain
rates rather than macroscopic quantities. One
obtains this form of the equations by dividing V,
V0, and Dc by the thickness of the sliding zone W.
This is a logical testable extension of the macroscopic equations allows modeling strain rate localization [Sleep et al., 2000]. The computed strain
rates are then mesoscale expectation functions
over enough grains for the gouge and enough time
to act as a continuum. The mesoscopic or grain
scale quantities then are shear strain rate e0  V/W,
reference strain rate e00  V0/W, and the intrinsic
strain eint  Dc/W. The friction equation (1) then
becomes




t P m0 a ln e0 =e00 b lny=ynorm ;

a=b

P0 e0

Pa=b
a=b

[ 8 ] The evolution equation for porosity then


becomes
@f
Ce e0 Ce e00 Pa=b

;

a=b
eint
@t
yeint P

8a

f 0 be0  C1 Pn exp f =Ce :

If the steady state sliding friction is independent of


the normal traction (3) and (5) imply that
ynorm

where f is a reference porosity and Ce is a


dimensionless material property. This equation
applies to gouges made mainly of hard grains like
quartz where the porosity is low enough (<10%)
that shear strain dilates rather than compacts. I
implicitly restrict discussion to processes occurring
at constant temperatures and effective normal
tractions appropriate for the seismogenic zone,
315 km depth.

The steady state value of the state variable in (4) is


Pa=b e00

yss



ff
;
y exp
Ce

where the first term implies that the dilational strain


rate is proportional to the shear strain rate and the
second term resembles power law compaction from
the normal traction P with the state variable y
representing the nonlinear effect of porosity on
compaction. In a simpler form the equation
becomes

and the evolution equation (2) becomes


@y
e0 Pa=b ye0
:
0 a=b 
eint
@t
eint P

especially when I later introduce the Ruina [1983]


evolution law. The state variable is then

P0

(It has not escaped me that a more complex


expression could represent the subtle variation of
friction with normal traction.) To this point, the
equations are mathematically equivalent to rate and
state friction with the Dieterich [1979] evolution
law and the Linker and Dieterich [1992] formulation for changes in normal traction.
[7] One would like to relate the state variable to a
physically measurable parameter. The porosity f is
a natural choice. I follow the formulation of Segall
and Rice [1995] as it leads to compact expressions,

8b

The first term indicates that the dilational rate is


linearly proportional to the shear strain rate with
the constant b. The second term represents
compaction with the constants collected into C1.
It has the reasonable properties that the compaction
rate increases as a power n of normal traction and
exponentially with porosity. Note that my use of
ynorm in (4) implies that the porosity in (8) and the
state variable in (4) cannot change suddenly when
normal traction changes. This is a reasonable
assumption or approximation.
[9] The steady state porosity in (8a) is
"
fss f  Ce ln

Pa=b e00
a=b

P0 e0

#
:

It is also illustrative to express the friction equation


(3), as a flow law for the shear strain rate.
Combining (3) and (6) yields
e0 e00

P
P0

a=a
b=a

1
t  m0 P
:
exp
y
aP

10

3 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

[11] A general expression for the thermally activated creep rate of a substance allows relating
terms [Sleep, 1997]

q
treal M
1 ;
e0real e0base exp
RT

Figure 2. Stresses on a plane in shear-traction,


normal-traction space. Frictional sliding occurs within
the shaded region where the ratio t/P is approximately
m0. Higher ratios cannot be reached at realizable sliding
velocities. The shear strain rate is immeasurably slow at
lower ratios, where gouge compacts.

The denominator in the exponential term of (10) is


small, as a is of the order of 0.01 and t/P is of the
order of 1. If one defines the constants m0, e00 , and
P0 so they refer to achievable values within a
simulated gouge or a slipping crustal fault zone,
t/P is approximately m0, a constant coefficient of
friction (Figure 2). The predicted sliding is
undetectably small for significantly lower ratios
and unreachably large for larger ratios. As a
practical matter, (10) and (3) need to give good
predictions only over this narrow range of ratios.

2.2. Friction as Creep Process


[10] To see how the behavior in (10) arises, I
consider the micromechanics of friction. It is
generally agreed that rate and state friction is a
thermally activated process [e.g., Berthoud et al.,
1999; Baumberger et al., 1999; Rice et al., 2001;
Nakatani, 2001; Nakatani and Scholz, 2004;
Beeler, 2004]. A simple form of the derivation
proceeds from real contact theory. Microscopic real
contacts with normal traction Preal exist along the
mesoscopic material contact. The rest of the mesoscopic surface is voids. The shear traction on the
contacts scales to the macroscopic values
treal

tPreal
:
P

11

The simple interpretation of a coefficient of friction


comes from treating treal as a yield stress, which
gives m0 = tyield/Preal.

12a

where e0base is a material constant, M is the


molecular volume of the rate limiting step, R is
the gas constant, and T is absolute temperature. In
the low-stress limit, this equation reduces to power
law creep with the exponent q, which can be seen
by taking the Taylor series of the exponential.
Indentation tests indicate that an amorphous or
highly disordered material forms at real contacts
[Goldsby et al., 2004]. This may indicate that a
linear material with q 1 is a better value than the
exponent for dislocation creep of 35.
[12] Here I am interested in the creep of hard
mineral grains at room to seismogenic temperatures. Shear creep can occur at a measurable rate
only at high stresses where the exponential term is
much greater than 1. Ignoring the 1 yields a
simpler expression:

qtreal M
:
e0real e0base exp
RT

12b

The microscopic strain rate in (12b) should be


proportional to the mesoscopic strain rate in (10). I
compare the term within the exponential in (12b)
with the term within the exponential in (10). (Note
the m0 term in (10) can be expressed as a
multiplicative constant outside the exponential.)
This yields using (11)
a

RT
:
qMPreal

13

The derivation is adequate to this point and an


analogous derivation follows for b [Nakatani and
Scholz, 2004]. One needs to explicitly consider the
micromechanical interaction of shear-driven creep
and compaction creep to understand evolution laws
and to get at the stability parameter b  a.

3. Ruina Evolution Law


[13] An alternative evolution law by Ruina [1983]
has the property that no healing occurs when
sliding is stopped. It has this form in my notation:
"
#
a=b
@y
e0 y
P0 e0 y
ln a=b 0 ;

@t
eint
P e0

14

4 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

Figure 3. The normalized time derivative of the state


variable after a sudden change in strain rate. The new
velocity is normalized to the previous strain rate, that is,
e0/e00 in terms of mesoscopic variables. The Ruina [1983]
law tends to zero for small velocities, while the
Dieterich [1979] curve tends to 1. The hybrid curve is
1/10 Dieterich [1979] curve plus 9/10 Ruina [1983]
curve. It is intended to represent laboratory situations
that approximately follow the Ruina [1983] law with
modest changes in strain rate and that heal somewhat
over long times during holds.

where the steady state is given by (5) and the


state variable is dimensionless. My form (14)
is mathematically equivalent to the Linker and
Dieterich [1992] relationship combined with
the Ruina [1983] law. With the shear-traction
equation (3) it predicts the same changes of shear
traction for changes in normal traction.
[14] Simulated gouge under low relative humidity
shows this behavior in the sense that friction does
not increase from its previous value when sliding
restarts after a hold [Frye and Marone, 2002].
However, the prediction of the porosity-state equation (7) and (14) that porosity does not change
during holds is not observed. Frye and Marones
[2002] gouge compacted during low-humidity
holds. I consider the physics of a nonlinear granular
material to show how the Ruina [1983] law arises. I
discuss the observed compaction in section 4.
[15] The different forms of the evolution law in
(14) and (4) matter greatly in situations relevant to
the earthquake cycle. First, sliding may slow down,
allowing healing to occur. I show this effect
without loss of generality in normalized form
where both the steady state-strain rate before sliding slows and the reference strain rate have values
of 1. (We can set the reference strain rate and
reference velocity for convenience with the restric-

tion they cannot be 0. The previous sliding velocity


is convenient in a laboratory experiment. The longterm geological sliding rate might be convenient
for an actual fault.) I let the reference normal
traction and the normal traction be equal. The
steady state value of the state variable y is then
1. Figure 3 shows the time derivative of the state
variable immediately after a sudden velocity
change. The Ruina [1983] curve is similar to the
Dieterich [1979] curve between its maximum at
normalized e0 = 1/e and its initial value of 1. That
is, the choice of evolution law does not matter
much for subtle velocity changes. At lower velocities, the laws are quite different. The Ruina [1983]
curve tends to 0, while the Dieterich [1979] curve
approaches 1.
[16] A second simple case involves the state variable when preseismic creep resumes following a
significant length of time after an earthquake.
Sliding during the earthquake leaves the state
variable at a small and unknown value. With the
Dieterich [1979] evolution law, the state variable
increases proportionally to Pa/b in (4) with time to
a large value where the initial small value is
irrelevant. Porosity decreases with the logarithm
of time in (8b) at constant P. When the Ruina
[1983] law is taken literally, neither the state
variable nor the porosity changes from their values
immediately after the earthquake until preseismic
creep restarts years later.

3.1. Compaction as Nonlinear Creep


[17] Equation (8) models compaction of the fault
zone as power law creep. The shear traction does
not enter into the final term representing compaction. However, the macroscopic shear traction and
the macroscopic normal traction are of the same
order as the coefficient of friction is of order 1. One
would expect that these tractions could interact
within a nonlinear substance. A molecule within
a grain cannot tell the macroscopic origin of the
microscopic stress field.
[18] To start generally, I represent the microscopic
strain rates as tensors. The deviatoric strain rate in
an isotropic viscous nonlinear material is [e.g.,
Schubert et al., 2001, pp. 248249]
e0ij

sij n1
s ;
htn1
ref

15

where e0 is the strain rate tensor, ij are tensor


indices, h is a material property with dimensions of
viscosity, tref is a reference stress, n is the exponent
5 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

of the power law rheology, sij is the deviatoric


p
stress tensor, and s 
sij sij is its second
invariant. Equation (15) applies within the mineral
grains that are incompressible in irreversible flow. I
am interested in the deformation of porous gouge.
Both the normal traction P that drives compaction
and the shear traction t lead to deviatoric strain
within the mineral grains. Both do macroscopic
work. The other components of the stress tensor
within the gouge depend on them. The invariant
governing creep thus involves both these stresses
[e.g., Wong et al., 1997]. The simplest form for the
macroscopic shear strain is then
e0

n1=2
t  2
t c2 P2
;
htn1
ref

16

where c is a dimensionless constant of order 1 that


takes into account that the gouge is a porous
substance and the dimensionless constant from the
different definition of the invariant in (15) and (16)
can be included in h. The compaction rate of the
gouge is similarly
f 0

n1=2
EP  2
t c2 P2
 be0 ;
n1
htref

17

where E is a dimensionless constant that represents


the different ductile compliance of the material in
compaction than in shear. The final term linearly
represents dilatancy as in (8b). It is an anisotropic
term arising from the grain geometry of the gouge.
Note that I have made the assumption that the
gouge is a thin tabular layer so that the rate of
approach of the walls of the gouge zone is the
negative rate of change of porosity times the gouge
zone thickness, f 0W.
[19] Strictly, an analogous term to the second righthand-side term of (17) should be included in (16).
It represents compaction creep driving shear creep.
That is,
e0

n1=2
t  2
t c2 P2
Wf 0 ;
n1
htref

18

where W is a constant with unknown sign of the


order of b. If I restrict the analysis to situations in
which frictional sliding is actually occurring
somewhere near steady state, then both terms in
(17) and f 0 are of the order of be0, where b is on the
order of a few percent. The cross term in (18) is
thus of the order b2e0 and can be ignored.
[20] Equations (16) and (17) represent the macroscopic effects of exponential creep at asperities
(over a limited range of real stresses) with a power
law rheology, before presenting an exponential
rheology in section 3.3. The exponent n is large

and the invariant dominates in (16) and (17). The


creep rate goes from very small to very large over a
limited range of the invariant. If one defines failure
as the occurrence of interestingly fast irreversible
creep, this feature is a simple form of the very
useful approximation of an elliptic failure envelope
in viscous porous materials [e.g., Wong et al.,
1997; Rudnicki, 2004].
[21] With forethought, I represent the compaction
rate as a function of the shear strain rate

EP
f e
b :
t
0

19

I wish to associate terms for the case where


frictional sliding is occurring and the friction
equation (3) is applicable. Using (3) for P/t yields
f 0 e0


 0



E
a
e
b
y
1  ln 0  ln
 e0 b; 20
e0
m0
m0
m0
ynorm

where I have used the Taylor series expression


1/(1 + x) 1  x to bring the second order terms into
the numerator. By assumption, b does not depend on
the velocity or the state variable. A steady state then
can be reached only if it cancels the constant term in
the bracket. That is, I associate b = E/m0.

3.2. Relationship of Nonlinear Creep to


Ruina Law
[ 22 ] Making this association, equation (20)
becomes
f 0 e0


 0



E a
e
b
y
:
ln 0 ln
e0
m0 m0
m0
ynorm

21

Continuing I express the left hand side of the Ruina


[1983] evolution law (14) in terms of the Segall
and Rice [1995] porosity relationship (7). Expanding the logarithm yields,
 0



Ce
e
y
;
f e
ln 0 ln
eint
e0
ynorm
0

22

where I use (6) to compact notation. It is intriguing


that (21) and (22) have similar forms. They become
homologous when a = b. Otherwise the two steady
states are not consistent. The implications of (21)
and (22) are in any case similar, reducing shear
traction not only decreases the shear strain rate as
expected, but also the compaction rate.
[23] Note that (22) cannot apply in the limit of
intact rock f = 0 . Then using (6) and (7) and letting
f be the steady state porosity at feasible e0 = e00 and
6 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

(Note that a cross term tmPm is possible if the


stresses locally drive flow in the same deviatoric
sense; then careful attention needs to be given to
signs.) The microscopic shear strain rate is
[Berthoud et al., 1999, equation (22)]
e0m g0 sinq expr=s;

25

where g0 is a material constant with dimensions of


strain rate and s is a material property with
dimensions of stress. It is related to RT/M in (12)
and is on the order of 100 MPa [Nakatani and
Scholz, 2004] as shown at the end of this section.
The corresponding expression for compaction is
fm0 g0

Figure 4. Microscopic stresses in parametric r, q


space. The red box indicates stresses that drive creep. Its
angular width is d. Lower values of the invariant r do
not produce significant creep; such stresses occur, for
example, within the interior of grains. Higher values
relax and are not present. The parameter l represents the
tendency for shear stress concentrations to persist longer
than normal stress concentrations.

P = P0, one gets f 0/e0 = f/eint, which is of the order


of 1 [Sleep et al., 2000]. This unlikely limit does
not occur in the Dieterich [1979] law (8), which
gives an acceptable description of the initial friction in intact rock as a function of normal traction
[Sleep, 1999b].

3.3. Micromechanics of Exponential Creep


[24] I continue the approach of the previous section
by explicitly considering the microscopic variation
of stress within the gouge and including exponential creep as in (12). This leads to compact expressions, and it is better to use the actual exponential
rheology rather than the approximate power law.
[ 25 ] It is convenient to express microscopic
stresses with parametric equations in q with analogy to the invariant in (16) (Figure 4),
tm  r sinq;

23a

cPm  r cosq:

23b

and

where the invariant is [Berthoud et al., 1999,


equation (20)]
r

q
t2m c2 Pm2 :

24

 
E
cosq expr=s  be0m ;
c

26

where a dilatancy term is included as in (17). The


strain rates go from extremely small to extremely
large over a small range in the invariant r, which is
(when considered less precisely) the yield stress or
real strength of the material on the order of a few
GPa for rocks. That is, the material is at a critical
state where many asperities are close to failure. As
already noted, the solid grains are incompressible
so that only microscopic deviatoric stresses within
the grains produce deviatoric creep. I have made
the simplifying assumption that this deviatoric
creep can be divided into creep that produces
mesoscopic shear strain and creep that produces
compaction.
[26] The instantaneous grain scale strain rate is
the integral of the actual instantaneous microscopic strain rates over a region in the gouge. I
apply this concept to special cases in sections 3.4
and 4. For now, I consider mesoscopic strain
rates that are spatial and brief time integrals of
the microscopic strain rates over the gouge. I
assume that the gouge is chaotic so that one can
only predict its statistical properties. This is
reasonable for simulated or real gouge with
irregular fragments of various sizes. I also assume that the grain and asperity sizes are small
enough that the finite thickness of the gouge
zone does not create order. My formulation is
thus not applicable to systems with high order
like aligned rods and sorted spheres capable of
rolling studied by Anthony and Marone [2005].
[27] I pose the problem statistically with integrals
over a weighting or probability density function g
in analogy with Figure 8 of Aharonov and Sparks
[2004]. Here I integrate formally over values of r
and q to get the expected values of mesoscopic
7 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

properties (Figure 4). The mesoscopic shear strain


rate over a finite time is formally,
e0

ZZ

rdr dqgq; re0m q; r;

27

where the dimensions of g are inverse stress


squared. The weighting function need only be
bounded for small values of r where the strain rate
is extremely small (Figure 4). It must go to zero
with r faster than the strain rate in (25) so that the
integral stays bounded above the yield stress.
Physically these conditions represent that one
expects to find few extreme stress asperities that
would relax quickly and that most of the deformation occurs near the yield stress. Equivalently, the
weighting function need only represent the distribution in the domain of stress components where
creep actually occurs. The q dependence needs to
represent that the microscopic stresses scale to their
mesoscopic and macroscopic quantities. A convenient form of (27) satisfies these inferences
e0 g0

Z
hrrdr

h
i
d
p sinq exp d2 q  q0  l2 dq
p
28

where h is the r -dependent product of the


weighting function and the invariant (dimensions
of inverse stress squared). Reiterating, the weighting function h is significant only for values of the
invariant stress r near the yield stress. Extremely
sluggish creep occurs at low values and rapid creep
and stress relaxation precludes high values. Otherwise we need not know the rheology in detail. The
dimensionless constant d is proportional to the
standard deviation of the Gaussian distribution
(Figure 4). I presume it is small so that microscopic
parameters scale to macroscopic ones. The other
parameters within the bracket are
tanq0 

t
;
cP

29

which represents the tendency of the ratio of the


microscopic stresses to scale with the mesoscopic
stresses. The parameter l represents the intuitive
tendency of asperities with high normal traction to
behave differently than those with high shear
traction (Figure 4). Well-aligned shear zones allow
an indefinite amount of slip, while compaction
closes pore space. In more detail, creep driven by
normal traction brings touching grains closer
together, increasing the microscopic contact area
and decreasing the microscopic stress. Creep
driven by shear traction parallel to a contact may

even decrease contact area. Thus a high shear


traction contact may accommodate more creep than
a high normal traction contact during its lifetime. I
define l so that is positive when high shear traction
contacts tend to persist longer than high normal
traction ones.
[28] As intended, I obtain simplification as r and q
parts of the integral separate. Taking the Taylor
series for sin(q) about q0 and the limits of the
rapidly convergent integral at 1 yields simple
expressions
e0 g0

Z
hrrdrsinq0 l cosq0

30

The expression for compaction rate is


g0 E
f
c
0

hrrdrcosq0  l sinq0
 be0 :

31

I assume that we are in the range of parameters


where frictional sliding occurs and apply the
friction law (3) for t/P terms from tan (q0) in
(29) so that I can associate terms. This yields that
the ratio of compaction rate to shear strain rate as in
(19) is

 0


f 0
E
a
e
b
y


1

ln
ln
e0
e00
m0
m0
m0
ynorm
 2


2
m c
b
l 0
m0 c

32

where small Taylor series terms are retained to first


order. That is, l2, a2, b2, al, bl, ab and higher
terms are ignored. The term in the brackets
multiplying l is positive so that the effect of l is
to retard compaction relative to shear as expected.
As with (20), the constant term represents
compaction balanced by dilatation b = E/m0. The
porosity change with this association is

 2


 0


e0 E a
e
b
y
m0 c2
:
l
f
ln 0 ln
e0
m0 c
m0 m0
m0
ynorm
0

33

Regrouping the terms yields an expression that


provides some insight into the mesoscopic and
macroscopic implications of the variable l,
f0


 0


 2

e0 E b
e
b
y
m c2
l 0
ln 0 ln
e
m0 c
m0 m0
m
ynorm
 00 
0
ab
e

ln 0 :
e0
m0

34

The first two terms in the bracket yield an


expression in the form of the Ruina [1983]
8 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

evolution law in (22) and its steady state if the


latter two terms cancel. That is,
 2

 0
m c2
ba
e
l 0
ln 0 :

m0 c
e0
m0

35

3.4. Changes in Normal Traction

Differentiating yields



@l
m20 c2
:
ba
c
@ ln e0

36

That is, l must change with strain rate to change


steady state properties. The parameter b  a is
related to the tendency of shear-traction concentrations to persist longer relative to normal-traction
concentrations with increasing slip rate. This and
the definition of l in (28) are simple kinematic
interpretations of rate and state friction.
[29] Continuing, the exponential terms in (25) and
(26) represent contact theory as does (12). I obtain
a from the effect of a sudden strain rate change.
The contact asperity stays the same (for a little
while) so the effect is to add microscopic shear
traction proportional to the mesoscopic change in
strain rate. The change in (3) or (10) is
@ lne0
1

:
@t
aP

37

Differentiating (25) and applying (24) yields


1=2
@ lne0 tm  2

:
t c2 Pm2
s m
@tm

38

Applying the first-order relationship that tm = m0Pm


to terms in the bracket and letting microscopic
quantities scale to mesoscopic ones in the bracket,
tm = tPm/P, yields
1=2
@ lne Pm  2

:
m c2
sP 0
@tm
0

39

A condition for (39) is that the assumed first-order


relationships hold. Comparing (37) and (39) yields
a

1=2
s  2
m0 c2
:
Pm

40

It is useful to compare the form of (40) with that of


the thermally activated creep expression in (13).
The microscopic pressure Pm corresponds to the
real normal traction when frictional sliding actually
occurs and
s

This allows approximate evaluation of s. The term


in the bracket is of order 1. For a laboratory
experiment on quartz, T 300 K, q 1
(for amorphous damaged material), and M 2 
105 m3 mol1. This yields 120 MPa.

1=2
RT  2
:
m c2
qM 0

41

[30] The friction equation (3) predicts that shear


traction changes instantly when normal traction
changes. This feature is impossible, as modeling
fault rupture between materials with different elastic constants with this property leads to a physically ill-posed problem [Ranjith and Rice, 2001].
That is, the Linker and Dieterich [1992] law
included in (3) applies to sudden changes where
stress concentrations have had time to self organize
at the asperity level and evolve the statistical
distribution implied by g in (27). Contact theory
provides that shear traction does not change when
normal traction instantly changes at constant sliding velocity. The microscopic contact area in (12)
does not instantly change so the real shear traction
does not change either. That is, the geometry of
asperities depends on the past history of normal
traction not its instantaneous value [Ranjith and
Rice, 2001]. In terms of (10), the pressure and state
variable terms represent the tendency of asperities to
persist in the lattice. They reflect a time-dependent
process and hence cannot change instantly.
[31] Perfettini et al. [2001] discuss how to formulate the evolution law so that shear traction does
not suddenly change in such a case. In my notation,
the state variable does not change instantaneously
and I need to modify the expression for ynorm in
(6). A convenient form is

lnynorm

Zt0
m
Pt0
wm  a

ln
b
P0
b
1

P
dt;
 expwt0  t
ln
P0

42

where P(t0) is the instantaneous normal traction,


the instantaneous coefficient of friction is m  t/P,
and w has units of 1/time. This expression has the
property that an instantaneous change in normal
traction does not instantaneously change shear
traction. If normal traction varies on times longer
than 1/w, it becomes the Linker and Dieterich
[1992] relationship in (6). This can be seen by
substituting P = P(t0).
[32] A bulky but more physical delay term would
involve strain rather than time. The quick response
9 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

from (42) and the longer-term response from the


evolution equation (4) or (14) separate if the quick
strain is much less than eint, 0.060.12 using the
values obtained by Sleep et al. [2000], who
accounted for strain rate localization. It has not
escaped me that this is similar to the porosity in the
gouge.
[33] A related effect is that the real area of asperity
contact immediately after a hold is higher than that
when friction reaches a peak after a small amount
of shear strain [Goldsby et al., 2004]. The contact
areas increase logarithmically with hold time during the hold as expected from an exponential
rheology, but a first increment of shear strain partly
disrupts these contacts without changing porosity.
The Segall and Rice [1995] relationship with the
friction law (3) is an approximation that applies
when enough sliding has occurred for an expectation function to apply.
[34] Qualitatively, there are two effects at real
contacts. First, creep from normal traction strengthens an individual contact and brings the grains
closer together. This convergence compacts the
gouge making more contact area elsewhere. Second, shear creep rearranges the contact where it
occurs and makes and breaks other contacts elsewhere on a rough surface. This effect occurs at
constant porosity and is distinct from dilatancy.
[35] Thus both shear and compaction strain need to
be included in a quantitative model of a gouge
where normal traction changes. For example, one
might expect that the Linker and Dieterich [1992]
parameter a to differ between a sudden increase in
normal traction where the gouge layer rapidly
compacts and a sudden decrease in normal traction
where rapid shear strain rearranges real contacts
and produces dilatancy.
[36] Measurements to resolve whether the parameter a depends on the sense of the change in
normal traction are inadequate to resolve the issue.
One complication is the failure of the patch (millimeter scale, Figure 1) normal traction on a
laboratory contact may not be the same as the
macroscopic nominal value everywhere on the
gouge surface [Sleep, 1999a]. A second complication involves strain rate localization and delocalization [Sleep et al., 2000].

creep whenever the stress components at asperities


scale to the macroscopic components as was assumed in the derivation of (32). In particular, the
macroscopic strain rate depends on the integral of
microscopic strain rate or more tractably on an
integral of the expectation function for microscopic
strain (28), (30), and (31). The rate and state stability
parameter b  a represents the tendency of shear
stress asperities to persist lower at high strain rates
relative to normal traction concentrations (36).
[38] My derivation involved little detailed physics
and lattice dynamics so it should have some
generality. In particular, I did not need to use
whether the macroscopically limiting region of
exponential creep was at the yield stress near
asperities between grains or that at crack tips as
assumed by Beeler [2004]. One can do much better
using the real tensor compliance properties of
grains, cracks tips, and gouge lattices. This computation seems possible but protracted. One might
obtain the probability density function, the gross
coefficient of friction m0, and the form of the
invariants from a feasible number of grains. One
would then integrate to get macroscopic properties.
For example, Aharonov and Sparks [2004] obtain a
probability density function for stress with a 24 
25 two-dimensional lattice of grains and a rheology
with elastic grains and frictional contacts.
[39] I now return to situations where the Ruina
[1983] laws fails to represent the gross behavior of
frictional sliding. First, the compaction rate in the
Dieterich [1979] evolution law is independent of
the shear traction so that healing occurs during
holds. As such behavior sometimes happens in the
laboratory, it is relevant to see how it arises in real
samples. Examination of (17) and (26) provides
some insight. As implied by the weighting function
g in (27), microscopic components of the stress
tensor and the strain rate tensor differ from the
mesoscopic tractions at asperities. An extreme case
is that the normal traction asperities do not have
correspondingly high shear traction concentrations
and high shear traction asperities do not have high
normal traction concentrations. The high normal
traction term thus dominates the microscopic invariant in (17) at the asperities that accommodate
compaction. Taking the first order term yields
f 0

4. Discussion and Conclusions


[37] As shown in section 3, the gross behavior of the
Ruina [1983] evolution law arises from exponential

Ecn1 Pn
;
htn1
ref

43

which is independent of the shear traction. The


contact theory in (13) or (40) still applies with an
effective yield pressure and a real shear traction.
10 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

[40] Returning to physics, I qualitatively compare


high humidity where the Dieterich [1979] law
appears to hold with low humidity where the Ruina
[1983] law appears to hold. High humidity results
in hydrated silica, a weak material at asperities
accommodating shear strain [Frye and Marone,
2002; Di Toro et al., 2004; Anthony and Marone,
2005; Hong and Marone, 2005]. The microstrain
for compaction may occur within the stronger
anhydrous silica of the grains. This creep maintains
real contact areas. The net effect is to separate
shear strain from compaction so that the Dieterich
[1979] law applies. Conversely, the Ruina [1983]
law applies at low humidity because the essentially
anhydrous grains are more homogeneous substance.

represent the evolution law as a sum of a small


Dieterich [1979] in (4) and a large Ruina [1983]
term in (14),

[41] More subtly, the Ruina [1983] evolution law


does not apply when the gouge has been under low
shear traction for a long time so that no frictional
sliding has occurred. The friction law (3) used in the
derivation then need not apply at astronomically
small shear strain rates. Normalizing the compaction rate to the ill-determined minute shear strain
rate in (22) and (33) is not warranted. The kinematic
difference between shear and compaction is the
opposite of that during active sliding. Small
amounts of shear strain relax stresses that are not
renewed by macroscopic sliding. Compaction can
continue at a finite rate until pore space vanishes.

[45] Finally, crustal faults at seismogenic depths


are obviously wet. I expect that healing and compaction occur over a long interseismic interval. The
Dieterich [1979] evolution law or a hybrid law
may provide a good approximation. One may have
to explicitly consider pressure solution.

[42] In addition, strain rate localization during


sliding lets Dieterich [1979] and Ruina [1983]
behavior coexist during a hold. The gouge surrounding a high strain zone is far from frictional
failure and can continue to compact as in (27),
while the gouge within the high strain rate zone is
near frictional failure.
[43] Experiments support the inference that gouge
can compact far from frictional failure. For example, Hagin et al. [2005] present data showing that
the Dieterich [1979] compaction law (8) applies to
sand under isotropic compaction e0 = 0, which is as
far from shear failure as one can get in an experiment. Power law creep as in (17) and (43) is a
reasonable representation of these data. Note that
continuing compaction is kinematically less likely
than continuing shear to regenerate high-stress
asperities.
[44] Situations intermediate between the Dieterich
[1979] and Ruina [1983] evolution laws are
expected, complicated, and observed in the laboratory [Boettcher and Marone, 2004]. As a practical matter for modeling real materials, one can

"
#
"
#
a=b
@y
e00 Pa=b ye0
e0 y
P0 e0 y


ln a=b 0 ;

a=b
eD
@t
eR
P e0
eD P0

44

where eD and eR are the intrinsic strains for the


Dieterich [1979] and Ruina [1983] terms, respectively. The time derivative of the state variable in
Figure 3 then reaches a small asymptote at small
strain rates rather than going all the way to zero.
This expression as given retains the steady state in
(5). Kato and Tullis [2001] proposed a different
form of a hybrid evolution law with grossly similar
implications.

Notation
a
b
c
C1
Ce
Dc
e0
E
f
0
m

f
fss
g
h
ij
M
n
P
P0
Pm
Preal
q

rate coefficient, dimensionless.


state coefficient, dimensionless.
constant in stress invariant, dimensionless.
grouped constant for healing term, stressn.
material property relating porosity and state
variable, dimensionless.
critical displacement, m.
strain rate tensor, s1.
constant for ductile compaction, dimensionless.
porosity, dimensionless.
microscopic compaction rate, s1.
steady state porosity, dimensionless.
weighting function, stress2.
weighting function depending on stress
invariant, stress2.
tensor indices, dimensionless.
molecular volume, m3 mol1.
power law exponent, dimensionless.
normal traction, N m2.
reference normal traction, N m2.
microscopic normal traction, N m2.
real normal traction on contact, N m2.
power law exponent at low stress, dimensionless.
11 of 13

Geochemistry
Geophysics
Geosystems

R
r
s
t
T
V
V0
W
a
b
g0
d

h
q
q0
eint
eD
eR
e0
e00
em0

e0base

e0real
f
y
ynorm
yss
l
m
m0
sij
t
tm
treal
tref
w
W

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

gas constant, J K1 mol1.


stress invariant, N m2.
stress scale for exponential creep, N m2.
time, s.
absolute temperature, K.
sliding velocity, m s1.
reference sliding velocity, m s1.
thickness of the sliding zone, m.
coefficient associated with changes in
normal traction, dimensionless.
dilatancy coefficient, dimensionless.
material constant for microscopic creep, s1.
parameter representing variation of ratio of
microscopic shear to normal traction,
dimensionless.
viscosity for nonlinear material, Pa s.
parametric angle for microscopic stress,
dimensionless.
parametric angle for macroscopic stress,
dimensionless.
intrinsic strain, dimensionless.
Dieterich term intrinsic strain, dimensionless.
Ruina term intrinsic strain, dimensionless.
strain rate, s1.
reference strain rate, s1.
microscopic strain rate, s1.
material constant for exponential creep,
s1.
real strain rate at contact, s1.
reference porosity, dimensionless.
state variable, dimensionless.
normalizing coefficient for state variable,
dimensionless.
steady state value of state variable, dimensionless.
parameter representing persistence of shear
stress concentrations, dimensionless.
coefficient of friction, dimensionless.
first-order coefficient of friction, dimensionless.
deviatoric stress tensor, N m2.
shear traction, N m2.
microscopic shear traction, N m2.
real shear traction on contact, N m2.
reference stress, N m2.
decay factor in exponential, s1.
constant for compaction-driven shear, dimensionless.

Acknowledgments
[46] This research was in part supported by NSF grant EAR0406658. This research was supported by the Southern
California Earthquake Center. SCEC is funded by NSF Cooperative Agreement EAR-0106924 and USGS Cooperative
Agreement 02HQAG0008. The SCEC contribution number
for this paper is 896. I thank Jim Rice, Chris Marone, and
Terry Tullis for discussions at the 2004 SCEC meeting and
helpful emails. David Sparks and an anonymous reviewer
provided numerous suggestions.

References
Aharonov, E., and D. Sparks (2004), Stick-slip motion in
simulated granular layers, J. Geophys. Res., 109, B09306,
doi:10.1029/2003JB002597.
Anthony, J. L., and C. Marone (2005), Influence of particle
characteristics on granular friction, J. Geophys. Res., 110,
B08409, doi:10.1029/2004JB003399.
Baumberger, T., P. Berthoud, and C. Caroli (1999), Physical
analysis of state- and rate-dependent friction law, II. Dynamic friction, Phys. Rev. B, 60(6), 3928 3939.
Beeler, N. M. (2004), Review of the physical basis of laboratory-derived relations for brittle failure and their implications
for earthquake occurrence and earthquake nucleation, Pure
Appl. Geophys., 161(9 10), 1853 1876.
Berthoud, P., T. Baumberger, C. GSell, and J.-M. Hiver
(1999), Physical analysis of the state- and rate-dependent
friction law:: Static friction, Phys. Rev. B., 59(22), 14,313
14,327.
Boettcher, M. S., and C. Marone (2004), Effects of normal
stress variation on the strength and stability of creeping
faults, J. Geophys. Res., 109, B03406, doi:10.1029/
2003JB002824.
Di Toro, G., D. L. Goldsby, and T. E. Tullis (2004), Friction
falls toward zero in quartz rock as slip velocity approaches
seismic rates, Nature, 427(6973), 436 439.
Dieterich, J. H. (1979), Modeling of rock friction: 1. Experimental results and constitutive equations, J. Geophys. Res.,
84(B5), 2161 2168.
Frye, K. M., and C. Marone (2002), Effect of humidity on
granular friction at room temperature, J. Geophys. Res.,
107(B11), 2309, doi:10.1029/2001JB000654.
Goldsby, D. L., A. Rar, G. M. Pharr, and T. E. Tullis (2004),
Nanoindentation creep of quartz, with implications for rateand state-variable friction laws relevant to earthquakes mechanics, J. Mater. Res., 19(1), 357 365.
Hagin, P. N. (2004), Application of viscoelastic, viscoplastic,
and rate-and-state friction constitutive laws to the deformation of unconsolidated sands, Ph.D. thesis, 129 pp., Stanford
Univ., Stanford, Calif.
Hong, T., and C. Marone (2005), Effects of normal stress
perturbations on the frictional properties of simulated faults,
Geochem. Geophys. Geosyst., 6, Q03012, doi:10.1029/
2004GC000821.
Kato, N., and T. E. Tullis (2001), A composite rate- and statedependent law for rock friction, Geophys. Res. Lett., 28(6),
1103 1106.
Linker, M. F., and J. H. Dieterich (1992), Effects of variable
normal traction on rock friction: Observations and constitutive equations, J. Geophys. Res., 97(B4), 4923 4940.
Nakatani, M. (2001), Conceptual and physical clarification
of rate and state friction: Frictional sliding as a thermally
12 of 13

Geochemistry
Geophysics
Geosystems

sleep: evolution laws for rate and state friction 10.1029/2005GC000991

activated rheology, J. Geophys. Res., 106(B7), 13,347


13,380.
Nakatani, M., and C. H. Scholz (2004), Frictional healing of
quartz gouge under hydrothermal conditions: 2. Quantitative
interpretation with a physical model, J. Geophys. Res., 109,
B07202, doi:10.1029/2003JB002938.
Perfettini, H., J. Schmittbuhl, J. R. Rice, and M. Cocco (2001),
Frictional response induced by time-dependent fluctuations
of the normal load, J. Geophys. Res., 106(B7), 13,455
13,472.
Ranjith, K., and J. R. Rice (2001), Slip dynamics at an interface between dissimilar materials, J. Mech. Phys. Solids,
49(2), 341 361.
Rice, J. R., N. Lapusta, and K. Ranjith (2001), Rate and
state dependent friction and the stability of sliding between
deformable solids, J. Mech. Phys. Solids, 49(9), 1865
1898.
Rudnicki, J. W. (2004), Shear and compaction band formation
on an elliptic yield cap, J. Geophys. Res., 109, B03402,
doi:10.1029/2003JB002633.
Ruina, A. (1983), Slip instability and state variable laws,
J. Geophys. Res., 88(B12), 10,359 10,370.

Schubert, G., D. L. Turcotte, and P. Olson (2001), Mantle


Convection in the Earth and Planets, 940 pp., Cambridge
Univ. Press, New York.
Segall, P., and J. R. Rice (1995), Dilatancy, compaction, and
slip-instability of a fluid-penetrated fault, J. Geophys. Res.,
100(B11), 22,155 22,171.
Sleep, N. H. (1997), Application of a unified rate and state
friction theory to the mechanics of fault zones with strain
localization, J. Geophys. Res., 102(B2), 2875 2895.
Sleep, N. H. (1999a), Effects of extrusion of fault gouge on
frictional sliding, J. Geophys. Res., 104(B10), 23,023
23,032.
Sleep, N. H. (1999b), Rate- and state-dependent friction of
intact rock and gouge, J. Geophys. Res., 104(B8), 17,847
17,855.
Sleep, N. H., E. Richardson, and C. Marone (2000), Physics of
strain localization in synthetic fault gouge, J. Geophys. Res.,
105(B11), 25,875 25,890.
Wong, T.-f., C. David, and W. Zhu (1997), The transition from
brittle faulting to cataclastic flow in porous sandstones: Mechanical deformation, J. Geophys. Res., 102(B2), 3009
3025.

13 of 13

S-ar putea să vă placă și