You are on page 1of 37

INTERNATIONAL JOURNAL FOR NUMERICAL AND ANALYTICAL METHODS IN GEOMECHANICS

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


Published online 3 July 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nag.535

Transient solution for a plane-strain fracture driven


by a shear-thinning, power-law fluid
D. I. Garagash*,y
Department of Civil and Environmental Engineering, Clarkson University, Potsdam, NY 13699-5710, U.S.A.

SUMMARY
This paper analyses the problem of a fluid-driven fracture propagating in an impermeable, linear elastic
rock with finite toughness. The fracture is driven by injection of an incompressible viscous fluid with
power-law rheology. The relation between the fracture opening and the internal fluid pressure and the
fracture propagation in mobile equilibrium are described by equations of linear elastic fracture mechanics
(LEFM), and the flow of fluid inside the fracture is governed by the lubrication theory. It is shown that for
shear-thinning fracturing fluids, the fracture propagation regime evolves in time from the toughness- to the
viscosity-dominated regime. In the former, dissipation in the viscous fluid flow is negligible compared to
the dissipation in extending the fracture in the rock, and in the later, the opposite holds. Corresponding
self-similar asymptotic solutions are given by the zero-viscosity and zero-toughness (J. Numer. Anal. Meth.
Geomech. 2002; 26:579604) solutions, respectively. A transient solution in terms of the crack length,
the fracture opening, and the net fluid pressure, which describes the fracture evolution from the earlytime (toughness-dominated) to the large-time (viscosity-dominated) asymptote is presented and some
of the implications for the practical range of parameters are discussed. Copyright # 2006 John Wiley &
Sons, Ltd.
Received 1 February 2005; Revised 1 February 2006; Accepted 1 April 2006
KEY WORDS:

hydraulic fracturing; self-similar solutions; non-Newtonian fluids; power-law fluids

1. INTRODUCTION
The problem of a fluid-driven fracture propagating in rock arises in hydraulic fracturing, a
technique widely used in the petroleum industry to enhance the recovery of hydrocarbons from
underground reservoirs [1], as well as in the magma transport in the Earths crust by means of
*Correspondence to: D. I. Garagash, Department of Civil and Environmental Engineering, Clarkson University,
8 Clarkson Ave., Potsdam, NY 13699-5710, U.S.A.
y
E-mail: garagash@clarkson.edu
Contract/grant sponsor: Donors of The Petroleum Research Fund, American Chemical Society; contract/grant number:
ACS-PRF 36729-G2

Copyright # 2006 John Wiley & Sons, Ltd.

1440

D. I. GARAGASH

magma-driven fractures [2, 3]. The conditions under which fluid-driven fractures propagate in
rock vary widely (i.e. rock parameters, in situ stress and fracturing fluid rheology) and are
usually not well defined. In addition, hardly any in situ measurements regarding the fracture
dimensions and shape exist. Consequently, mathematical and numerical modelling of fluiddriven fractures have attracted numerous contributions, see, e.g. References [49] for some early
efforts. The main objective of the theoretical modelling is to predict the evolution of the fluid
pressure pf ; the fracture opening w and the size for given rock properties, fluid rheology and
injection rate under assumptions of simplified fracture geometry.
The modelling of hydraulic fracture requires consideration of both fluid and solid mechanics:
on one hand, the lubrication equation to characterize the flow of fluid in the fracture; and on the
other, the elasticity equations to describe the deformation and propagation of the fracture. The
solution of these models and their parametric analysis are notoriously difficult to develop
because of the strong non-linear coupling between the lubrication and elasticity equations and
the non-local character of the elastic response of the fracture. Non-locality implies that the
crack opening w at one location along the fracture depends on the fluid pressure pf at another
location. The non-linearity stems from the dependence of the fracture conductivity on a certain
power of the crack opening. Non-linearity and non-locality yield a complex solution structure
involving processes at a very small lengthscale near the fracture tip, which, nevertheless, control
the global response of a fluid-driven fracture [1015]. Recent rigorous treatment of hydraulic
fracture models of idealized plane-strain and axisymmetric geometries in the case of a
Newtonian fracturing fluid (see, e.g. References [1521]) has allowed to identify different
limiting propagation regimes, some of the corresponding asymptotic solutions and respective
behaviour of these solutions near the fracture tip. In the particular case of a fracture
propagation in an impermeable rock, the two limiting regimes can be identified with the
dominance of one or the other of the two energy dissipation mechanisms corresponding to
extending the fracture in the rock and to flow of viscous fluid in the fracture, respectively. In the
viscosity-dominated regime, dissipation in extending the fracture in the rock is negligible
compared to the dissipation in the viscous fluid flow, and in the toughness-dominated regime, the
opposite holds. Corresponding self-similar asymptotic solutions for Newtonian fluid case are
given by the zero-toughness [17, 22] and zero-viscosity [21] solutions, respectively. In the case of
a plane-strain fracture and a constant fluid injection rate, the propagation regime is controlled
by a single non-dimensional time-independent constant with the meaning of either dimensionless toughness or viscosity.
In this paper, we extend the analysis of the propagation regimes of a plane-strain hydraulic
fracture to the more general case of a fracturing fluid with a power-law rheology. This extension
is motivated by the recognition that fracturing fluids in petroleum and geophysical applications
often exhibit non-Newtonian (non-linear) behaviour [2325]. A class of non-Newtonian fluids
can be defined in the context of unidirectional laminar flow via the following rheological law,
which relates the shear stress in the fluid t to a certain power of the shear strain rate g :
t M g n

where n is the power-law exponent (also called fluid behaviour index) and M is the consistency
index. For Newtonian fluids, n 1; and M m: Most of the fracturing fluids used in the
petroleum industry are characterized by relation (1) in a range of shear rate g with 04n51: This
range of n corresponds to the so-called shear-thinning behaviour, as the apparent fluid viscosity
mn M g n$1 is a decreasing function of the shearing rate.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1441

We proceed to show that the fracture propagation regime in the case of a shear-thinning fluid
is not fixed in time as for the case of Newtonian fluid, but rather evolves from the early-time
toughness-dominated regime to the large-time viscosity-dominated regime, since the governing
dimensionless viscosity parameter is an increasing function of time for any 04n51: The endof-time, zero-toughness solution has been investigated by Adachi and Detournay [22]. The
beginning-of-time zero-viscosity solution (see Reference [21] for Newtonian fluid case)
corresponds to the classical solution of a uniformly pressurized Griffiths crack with a special
constraint on the fracture volume. The transient solution which describes evolution of the crack
length, the fracture opening, and the net fluid pressure from the zero-viscosity to the zerotoughness limiting solution is computed herein numerically. The numerical technique is based
on the series expansion of the normalized solution for the opening and the net pressure over a
class of base functions in the spatial co-ordinate [15, 22] which satisfy the elasticity singular
integral equation and conform with the proper tip asymptote. The evolution of the solution in
time corresponds to that of the series coefficients, governed by the set of ordinary differential
equations (ODEs) resulting from the spatial discretization of the fluid flow equation. Finally,
some implications of the transient solution for a practically relevant parametric range in
hydraulic fracturing are discussed.
2. MATHEMATICAL MODEL
Consider a finite two-dimensional hydraulic fracture of half-length t propagating in an
impermeable linear elastic medium characterized by Youngs modulus E; Poissons ratio n; and
toughness KIc : An incompressible fluid with power-law rheology, behaviour index n and
consistency index M; is injected at the centre of the fracture at a constant rate Qo : The crack is
loaded by the internal fluid pressure pf x; t and by the far-field confining stress so (positive in
compression), see Figure 1. The fracture is assumed to be in mobile equilibrium at all times and
its propagation is quasi-static. We will look for the solution of the fluid-driven fracture problem:
net pressure pf x; t $ so ; fracture opening wx; t; and fracture half-length t; where x is the
position along the crack and t is time. The fracture is assumed to be fully fluid-filled at all times,
i.e. there is no lag between the fracture and fluid front. Validity of this assumption is discussed in
the next section.
In the following we formulate the governing equations over the half of the crack, 04x4;
accounting for the crack symmetry with respect to its centre. Additional discussion of governing
equations can be found in, e.g. References [15, 21].
The fluid flow inside the fluid-filled crack is described by equations of lubrication theory [26],
namely, by the integral form of the continuity equation:
Z
@
w dx q
2
@t x
where q is the fluid flow rate per unit (out-of-plane) width of the crack; and by Poiseuilles
law [23]:

w2n1 @p
3
M 0 @x
where M 0 fM is a modified consistency index with f 2n1 2n 1n =nn being a numerical
factor. Alternatively, Poiseuilles law for a non-Newtonian fluid (3) can be written in the form
q j q jn$1 $

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1442

D. I. GARAGASH

Figure 1. Sketch of a plane-strain fluid-driven fracture.

for an equivalent Newtonian fluid


q$

w3 @p
;
m0 @x

m0 M 0

w2
jqj

"1$n

where the effective viscosity parameter m0 is given up to a numerical factor by the apparent
viscosity mn M g n$1 evaluated at the (crack) channel wall: m0 4n$1 2n 1mnwall (see
Appendix A for the details). For a Newtonian fluid, n 1; the effective viscosity m0 and the
effective consistency index M 0 coincide and equal, up to a numerical factor, to a constant
Newtonian viscosity m; m0 M 0 12m:
Global fluid continuity requires the injected fluid volume per unit crack width, Qo t; to be
equal to the fracture volume (zero fluid lag and no leak-off),
Z
2
w dx Qo t
5
0

In view of (2), the latter is equivalent to the boundary condition for the fluid flux at the inlet:
Qo
6
qx 0; t
2
The deformation of the solid and the fracture propagation criterion are prescribed by
equations of linear elastic fracture mechanics as follows. The net pressure in the crack p
pf $ so is related to the dislocation density @w=@x via the Cauchy singular integral [27], which
under provision of crack symmetry with respect to the origin can be written as
Z
E 0 @wx0 ; t x0 dx0
px; t $
7
@x0 x02 $ x2
2p 0
The inverse of the above integral relation providing the crack opening w in relation to the net
pressure p pf $ so is given by [28]
$pffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi$
"
Z !
$ 1 $ x2 1 $ x02 $
4
x x0
$
$
0
0
0
8
wx; t
G ;
px ; t dx ; Gx; x ln$pffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi$
0
$ 1 $ x2 $ 1 $ x02 $
pE 0

The criterion of fracture propagation in mobile equilibrium, which requires the stress-intensity
factor to be equal to the material toughness, KI KIc ; is expressed here in the form of the tip

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1443

asymptote of the crack opening [29]


w

K0
$ x1=2
E0

$ x{

In the above formulation, the three material parameters E 0 ; M 0 ; and K 0 with the meaning of
the effective elastic modulus, fluid consistency index, and rock toughness, respectively, are
defined as follows:
! "1=2
E
2
0
0
0
E
; M fM; K 4
KIc
10
1 $ n2
p
Equations (2)(9) fully define the fracture length t; the opening wx; t; and the net pressure
px; t as functions of the parameter set (10), the fluid behaviour index n; and the injection
rate Qo :

3. BEHAVIOUR NEAR THE CRACK TIP


In accordance with linear elastic fracture mechanics (LEFM) theory, the opening asymptotically
varies as the square root of the distance from the crack tip for any non-zero value of rock
toughness K 0 ; (9). The asymptotic behaviour of the net pressure gradient for the non-zero
toughness can be deduced from the analysis of the leading term in the lubrication equation (2)
with (3) and (9) for $ x{: The corresponding net pressure tip asymptote is then given
8
2
>
>
! 0 "1n
$ x1$n=2 ; n=1
<
E
1
$
n
n
0
$ x{
11
p constt M
vtip
> %
x&
K0
>
: ln 1 $ ;
n1

where constt is an integration constant (the part of the overall solution) and vtip d=dt is
the crack tip velocity. Consequently, the net pressure is always finite at the fracture tip for shearthinning fluids n51; and has a negative (logarithmic) singularity for Newtonian fluids n 1
and a negative algebraic singularity for fluids with n > 1: Since the fluid cannot sustain large
enough tension, cavitation has to necessarily occur near the tip of a fracture driven by a
Newtonian fluid or a fluid with n > 1; thus, resulting in a lag filled by fracturing fluid vapours
between the fracture front and the fluid front. For shear-thinning fluids n51; however, the tip
net pressure is finite and cavitation does not necessarily take place unless a finite value of fluid
pressure at the tip drops below the cavitation pressure. Given that the cavitation pressure is
typically negligible compared to the confining stress so ; the cavitation condition for the net
pressure implies p4 $ so : Since an unknown constant term (a function of the overall solution)
enters the pressure asymptote (11), the latter cavitation condition cannot be evaluated unless the
complete solution is known. However, it is interesting to note that the likeliness of this
condition to be satisfied decreases with either the increase of confining stress so or with the
decrease of the fluid behaviour index n (result of the weakening of the pressure gradient
singularity at the crack tip with decreasing n).
The above considerations, asymptotic tip boundary layer analysis under conditions of
vanishing effects of toughness (see below and in Appendix B), as well as the actual solution of
the problem discussed in Section 7 provide justification of the assumption of the zero fluid lag
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1444

D. I. GARAGASH

(or of a fully fluid-filled crack) employed in this analysis for the case of a shear-thinning fluid. It
is also worthwhile to point out that in the case of a fracture driven by a Newtonian fluid, even
when the lag is non-zero, it becomes vanishingly small fraction of the fracture length for long
enough fractures or large enough confining stress [12, 30].
In the case of zero rock toughness (e.g. a hydraulic fracture propagating along a preexisting
discontinuity), the coupling between elastic response (8) and the fluid flow in fracture (2)(3)
yields the tip asymptotic behaviour in the form [10]
K0 0 :

w W& $ x2=2n ;

p $P& $ x$n=2n

$ x{

12

with the coefficients W& and P& are given in the above reference as a function of the crack tip
velocity and material parameters. Since 2=n 2 > 1=2 for shear-thinning fluids, the viscosity
opening asymptote 12a is weaker than the LEFM one, confirming the zero toughness
assumption. One can expect that the zero-toughness solution (constructed for a power-law fluid
in Reference [22]) which possesses tip asymptote (12) has to be a valid approximation of the
solution with non-zero small enough values of toughness K 0 : The singular change in the tip
asymptote from the K 0 0 case, (12), to the, however, small but non-zero toughness LEFM
case, (9) and (11), therefore implies that for small values of toughness there exists a boundary
layer at the fracture tips in which the toughness effects are localized while the solution away
from the fracture tip is given by the zero-toughness solution. This boundary layer solution is
matching with the zero-toughness solution away from the fracture tips over an intermediate tip
lengthscale where both solutions posses zero-toughness asymptote (12), see Figure 2. The small
toughness solution with the structure described above has been explicitly constructed in the case
of Newtonian fluid n 1 in Reference [19]. The expression for the thickness of the tip
22n=2$n
#m
boundary layer, # K
; can be obtained from the direct application of the boundary
layer methodology [19] to the generalized governing equations for a power-law fluid case, given
# m is the dimensionless tip toughness parameter defined by (B8) in terms of the
in Appendix B. K
crack tip velocity, the global crack lengthscale ; and the set of material parameters, which
characterizes the smallness of the toughness effect compared to the one of the viscosity on the
global solution.

Figure 2. Sketch of the structure of the small toughness solution for fluid index n52: (The extent of
the near-tip boundary layer relative to the crack length is greatly exaggerated.) The near-fracture-tip
boundary layer provides transition between the toughness asymptote (9) at the immediate vicinity of
the fracture tip and the viscosity asymptote (12) at intermediate distances from the tip, where it is
matched with the zero toughness outer solution valid on the lengthscale of the fracture. (Variable x#
denotes distance from the fracture tip, $ x:)
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1445

Even though outside of this papers focus on shear-thinning fluids, it is interesting to note that
the structure of the tip boundary layer (Figure 2) in the small toughness solution becomes
invalid for shear-thickening fluids with sufficiently high index n > 2: Indeed, in this case the
viscosity asymptote for the opening 12a is stronger than the toughness one (9), and the
# m ! 0: This is an indication that the small toughness
boundary layer thickness diverges as K
solution does not have a tip boundary layer (i.e. it constitutes a regular perturbation of the zerotoughness solution). On the other hand, in the small viscosity (large toughness) case for n > 2;
the boundary layer is expected to accommodate the transition between the viscosity tip
asymptote valid immediately near the tip and the toughness asymptote at intermediate distances
from the tip, where it is matched with the zero-viscosity outer solution (classical solution for a
uniformly pressurized Griffiths crack).
In this paper, we focus on the finite toughness solution which makes use of the LEFM tip
asymptote specified by (9) and (11). In the view of the discussion above, this solution cannot be
used when toughness is zero or sufficiently small. It will be shown, however, that the developed
finite toughness solution for small enough values of toughness does approach the zerotoughness solution [22], and, therefore, together with the latter can be used to map the complete
range of toughness.
4. SCALING
Two physically meaningful scalings can be identified with conditions when the effect of the
material toughness K 0 on a fluid-driven fracture is large compared to the effect of fluid viscosity
m0 and when the opposite holds, respectively. Respective scalings can be dubbed toughness and
viscosity scalings, as the former is independent of the fluid consistency index M 0 (related to the
apparent viscosity) and the latter is independent of K 0 : The scaling methodology and the above
two scalings can be found in Reference [13] in the case of a Newtonian fluid. Furthermore, the
viscosity scaling for a power-law non-Newtonian fluid has been established in Reference [22]. In
the following we adapt the scaling methodology [13] to the case of a non-Newtonian fluid to
obtain both the viscosity and the toughness scalings by introducing the concept of an
equivalent Newtonian fluid with time-dependent characteristic viscosity.
Let us define the scaled co-ordinate x x=t 04x41 and the dimensionless field
variables: opening O; net pressure P; flux C; effective viscosity M; and the crack half-length g in
the following general form:
wx; t etLtOx; t

13a

px; t etE 0 Px; t

13b

qx; t etL2 tt$1 Cx; t

13c

m0 x; t m0 tMx; t

13d

t Ltgt

13e

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1446

D. I. GARAGASH

The above expressions prescribe a particular scaling of a hydraulic fracture for a given choice of
the fracture lengthscale L; the dimensionless parameter e (scaling the net pressure as a fraction
of the modulus E 0 ), and the characteristic viscosity m0 : Based on the global continuity (5), the
definition 4b of the effective viscosity m0 ; and requiring that all dimensionless field variables
(O; P; etc.) are O1; parameters e and m0 can be related to a lengthscale L:
!
"
2 1$n
Qo t
1$n
0
0
0 Qo t
M
14
e 2 ; m M et
L
L2
Thus, a particular fracture scaling (13a)(13e) with (14) can be specified by a single parameter,
the fracture lengthscale L:
4.1. Normalized equations
Introducing an alternative scaling for the opening and flux as
C
% O; C
%
O
g
g2

15

the governing equations (2)(9) and (11) in scaling (13a)(13e) with (14) can be reduced to the
following universal form.
*

Local fluid balance:


!
1 2G

"Z 1
!
"
@
d ln g
%O dx d GxO
% C
%
G
@ ln t x
d ln t

where d dln L=dln t is a constant when the fracture lengthscale is a power law of time,
L ' td :
Fluid flux:
% 2n1 @P
O
% jC
% jn$1 $
17
C
M @x
or its equivalent expression for an equivalent Newtonian fluid:
! 2 "1$n
% 3 @P
%
O
O
% $
; M
C
%
MM @x
j Cj

Global fluid balance:


1

2g2

19

% dx
O

Elasticity equation:
1
%
Px; t LfOgx;
t $
2p

Copyright # 2006 John Wiley & Sons, Ltd.

18

or its equivalent expression for the fluid flux at the inlet


1
% 0; t
Cx
2g2
*

16

1
0

% 0 ; t x0 dx0
@Ox
@x0 x02 $ x2

20

21

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1447

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

or in its inverse form


4
%
Ox;
t L$1 fPgx; t
p
*

Gx; x0 Px0 ; t dx0

22

Propagation condition and corresponding pressure gradient asymptote:


% Kg$1=2 1 $ x1=2 ; @P $A0 1 $ x$1n=2
1 $ x{1 : O
@x

23

where
A0 MK$1$n g1n=2 d Gn

24

The above set of normalized equations (16)(24) governs the normalized solution Fx; t
% P; C;
% gg which depends on the two dimensionless groups associated with viscosity and
fO;
toughness, respectively,
m0
K0
; K 0 1=2
0
3
tE e
eE L
0
where e and m are the functions (14) of a lengthscale L:
M

25

4.2. Viscosity and toughness scalings


For a particular scaling, lengthscale Lt can be specified by setting one of the dimensionless
combinations, (25), to unity, while the remaining dimensionless combination will constitute a
single dimensionless parameter governing the normalized solution of (16)(24). By setting either
M 1 or K 1 one can obtain the viscosity or the toughness scaling characterized by the
problem lengthscale L independent of the toughness K 0 or the characteristic viscosity m0 ;
respectively.
Resulting lengthscale Lt; the characteristic viscosity m0 t; and the dimensionless parameter
et in the toughness K and the viscosity M scalings and corresponding expressions for the
viscosity M and toughness K dimensionless groups are listed in Table I. Naturally, in the
toughness-scaling the solution scales (via scaling parameters e and L) with the material
toughness and the corresponding dimensionless toughness is unity; and in the viscosity-scaling
solution scales with the fluid viscosity and the corresponding dimensionless viscosity is unity.
The remaining non-trivial parameter (viscosity M in toughness-scaling, or toughness K in

Table I. Quantities corresponding to the viscosity (M) and the toughness (K) scalings.
m0

Scaling
K
M

M0

K 02 t
1=2
E 02 Qo

M 03=2n

L
!21$n=3

% t &21$n=2n
E 01=2

! 0
"
E Qo t 2=3
K0
1=2

E 01=6 Qo t2=3
m01=6

Copyright # 2006 John Wiley & Sons, Ltd.

K 04

E 04 Q
!

t
"1=3

m0
E0t

M
"1=3

Mk

E 03 Qo m0
K 04
1

1
Km

K0
1=4

E 03=4 Qo m01=4

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1448

D. I. GARAGASH

viscosity-scaling) governs the solution in the respective scaling. Corresponding indices k


and m ought to be added to all quantities (except for time t and space x) in the
normalized governing equations (16)(24) to designate equations in the two respective
scalings. (The time-exponent d of the fracture lengthscale, L ' td ; appearing in the fluid
balance equation (16) takes on values dk 2=3 and dm 1 n=2 n in the corresponding
scalings.)
Remarkably, the form of the two scalings for a power-law, non-Newtonian fluid presented in
Table I in terms of the corresponding characteristic time-dependent viscosities m0k;m t is
equivalent to the form of the respective scalings for a Newtonian fluid [13] for which m0k;m 12m
and m is a (constant) Newtonian viscosity. The expressions from Table I for the fracture
lengthscale L and dimensionless groups M and K in terms of the characteristic time-dependent
viscosities are expanded in Appendix C (Table CI) to reveal their explicit time-dependence. We
# m ; arising from the considerations
acknowledge that the dimensionless toughness parameter K
(of Section 3 and Appendix B) of the LEFM boundary layer near the front of an arbitrary
fracture propagating under conditions of small toughness, is identical up to O1 constant factor
to the dimensionless toughness Km for the plane-strain fracture (Table CI). In order to show the
latter equivalence of the these two measures of the effects of toughness on the fluid-driven
# m ; (B9),
fracture propagation, one need to set the global fracture lengthscale in the definition K
to the viscosity lengthscale Lm (Table CI) and use corresponding expression for the crack tip
m:
velocity vtip ' L
We further note that the solutions in the two scalings are simply interrelated via
!
"
! 0 "1=1$n
%m C
%m
gm Lk
Pm O
Mm 1=1$n
mk
ek

K$2=3
;

K4=3
26
m
m
%k
%k
gk L m
Pk
Mk
m0m
em
C
O
and
$4=3n2
Mk Km

27

The time-dependence of the governing dimensionless parameter (Mk ' t2=31$n or


Km ' t$1=21$n=2n ) is defined by that of the characteristic viscosity m0 in the corresponding
scaling (Table CI). Consequently, the evolution terms of the form t @(=@t in the local fluid
balance (16) can be expressed in terms of derivatives in either dimensionless viscosity Mk or
toughness Km with the help of the following identities:
@
2
@
11$n
@
1 $ n
$
28
@ ln t 3
@ ln Mk
2 2 n @ ln Km
% P; C;
% gg for
We then observe that the time-dependence of the normalized solution Fx; t fO;
the opening, pressure, flux and crack length comes only through the dimensionless viscosity or
toughness parameter, such that Fx; t Fk x; Mk and Fx; t Fm x; Km in the respective
scalings.
4.3. Special fluid cases (n 0 and n 1)
The two cases of fluid behaviour corresponding to perfectly plastic fluid n 0 and Newtonian
fluid n 1 correspond to the special cases where the solution-dependence on time is known in
advance.
In the Newtonian fluid case n 1; Mk and Km are time-independent constants, Table I,
thus, the evolutionary terms are zero, see (28), and the lubrication equations (16)(17) reduce to
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1449

a self-similar form
Z

% 3 dP
O
% dx dxO
% C
% $
O
M dx

G 0

29

The corresponding normalized solution is, thus, self-similar [15] (i.e. time-independent in either
scaling), and the dependence on time of dimensional opening, pressure and crack length is given
by that of the scaling parameter e and lengthscale L; Table I [13].
In the case of perfectly plastic fluid n 0; lubrication equations (16)(17) reduce to [22]
1$

% dP
O
M dx

30

And, even though, Mk ' t2=3 and Km ' t$1=4 are time-dependent, the reduced equations (30),
(19)(23) are no longer evolutionary since they depend on time as a parameter only.

5. PROPAGATION REGIMES
% P; C;
% gg of Equations (16)(24) in either M- or K-scaling
Normalized solution F fO;
depends on normalized position along the fracture x and a single governing dimensionless
parameter Mk or Km ; respectively. Consequently, the propagation regimes of a hydraulic
fracture can be completely defined in terms of one of these two parameters. Similar to the
analysis for the fracture driven by Newtonian fluid [13, 1721, 31, 32], the two limiting
parametric regimes corresponding to either Mk {1 (toughness-dominated regime) or Km
{1 (viscosity-dominated regime) can be identified with the situation where either the effect
of the fluid viscosity or the effect of material toughness on the hydraulic fracture
propagation is negligible, respectively. The solutions in these regimes can be approximated
by the zero-viscosity Mk 0 solution [21] or the zero-toughness (Km 0) solution [22],
respectively. These normalized limiting solutions are self-similar (i.e. time-independent) in
the respective scalings, such that the time-dependence of the dimensional solution is given
by that of the scaling parameters (Table I). Thus, according to the scaling discussion of the
previous section, the hydraulic fracture propagation in time driven by a non-Newtonian
fluid in either the toughness-dominated (early-time) or the viscosity-dominated (largetime) regime can be associated with the propagation of a fracture driven by an
equivalent Newtonian fluid with appropriate time-dependent viscosity, m0k t or m0m t;
respectively.
In view of Table I, Mk and Km are increasing and decreasing power-laws of time for shearthinning fluids n51; respectively. Thus, hydraulic fractures evolve from the toughnessdominated regime (with the zero-viscosity solution) at early time to the viscosity-dominated
regime (with the zero-toughness solution) at large time. This evolution can be tracked in either
of the two dimensionless parameters. The goal of this paper is to construct the transient solution
for the hydraulic fracture evolution between these two limits. Thus, first, we discuss the limiting
self-similar states of the transient solution, i.e. the large-time zero-toughness limit, the early-time
zero-viscosity limit, and, secondly, the small viscosity correction to the latter. The small (but
non-zero) viscosity solution will provide the starting point in time for the numerical transient
solution.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1450

D. I. GARAGASH

5.1. Zero-toughness self-similar solution [22]


In the limit of large (infinite) propagation time, the toughness parameter Km vanishes, and the
governing system of equations in viscosity-scaling reduce to a self-similar form
!
"
Z 1
Z 1
dPm0 1=n
$2
% m0 $O
% 2n1
% m0 dx 1 n xO
% m0 dx
;
g

2
31
O
O
m0
m0
2n
dx
x
0
% m0 x L$1 fPm0 gx;
O

% m0 x 0
lim 1 $ x$1=2 O

x!1

32

Equations (31) are the reduction of the lubrication equation and global fluid balance, and
Equations (32) are the elasticity equation and propagation criterion under condition of zero
toughness, respectively. The latter, 32b ; implies that opening has the asymptote at the tip other
than the LEFM type 23a : Indeed, the appropriate asymptote corresponds to the zero energy
dissipation in fracturing the solid and solely determined by the viscous dissipation in the fluid
flow. This asymptote, which follows from the asymptotic considerations for the lubrication
equation 31a and elasticity 32a ; is given by [10, 22]
8
> 1
1 $ xa$1 ; n=0
aa $ 1 cot pa <
a
%
) a$1
1 $ x{1 : Om0 Bm0 1 $ x ; Pm0 Bm0
33
>
4
: ln1 $ x;
n0

where

2
;
a
n2

Bm0

(a=2
!
" '
1 n 1$a 22 n2
tan$pa

2n
n

34

In the opening near-tip power law 33a ; the exponent a varies from 1 (wedge-like tip) at n 0 to
2=3 at n 1:
% m0 x; Pm0 x; gm0 g governed by Equations
The zero-toughness self-similar solution fO
(31)(32) with the near-tip asymptote (33)(34) has been derived by Adachi and Detournay
[22] using the method of the series expansion over the class of base functions satisfying elasticity
(32) and possessing the correct tip asymptote (33). The coefficients in the series expansion are
obtained by minimizing the quadratic error in the lubrication equation 31a : For an example of
their solution for fluids with n 0; n 0:5; and n 1; see the dimensionless net pressure and
opening in the viscosity scaling shown by dashed lines on Figures 4a0 c0 and 5a0 c0 ;
respectively.

5.2. Zero-viscosity self-similar solution


In the limit of infinitesimally small propagation time, the viscosity parameter Mk vanishes, and
the governing system of equations in toughness-scaling reduce to a self-similar form
Z 1
dPk0
$2
% 2n1
% k0 dx
0 $O
;
g

2
35
O
k0
k0
dx
0
% k0 x L$1 fPk0 gx;
O
Copyright # 2006 John Wiley & Sons, Ltd.

$1=2

% k0 x gk0 1 $ x$1=2
1 $ x{1 : O

36

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1451

The zero-viscosity self-similar solution of the above equations corresponds to a classical


solution of a Griffiths crack with the elliptic shape and uniform net pressure along the
crack [21]
1=3 qffiffiffiffiffiffiffiffiffiffiffiffiffi
p1=3
2
% k0 p
; O
37
Pk0
1 $ x2 ; gk0 2=3
8
2
p
5.3. Small-viscosity solution
The small-viscosity correction to the zero-viscosity solution (37) for a fracture driven by a
power-law fluid can be obtained by a direct adoption of the approach used in References [20, 21]
for the case of a Newtonian fluid. In the following, we briefly summarize the approach and the
results. Utilizing the toughness scaling, the small-viscosity solution Mk {1 can be sought in
terms of the series, Fk0 Fk0 Mk Fk1 ( ( ( ; as follows:
Pk Pk0 x Mk Pk1 x ( ( (

38a

%kO
% k0 x Mk O
% k1 x ( ( (
O

38b

gk gk0 Mk gk1 ( ( (

38c

where the first term Fk0 in the expansion corresponds to the zero-viscosity solution (37) and the
second term Mk Fk1 corresponds to the next-order viscosity correction. Substitution of (38a)
(38c) in the governing equations in the toughness scaling (16)(24) with Table Ik and 28a
while retaining terms of O1; provides governing equations (35)(36) for the zero-viscosity
solution (37). Similarly, retaining terms of OMk provides governing equation for the nextorder term Fk1 in expansion (38a)(38c), which solution yields
Pk1 Pk1 0 DPk1 x

39

qffiffiffiffiffiffiffiffiffiffiffiffiffi
% k1 4Pk1 0 1 $ x2 L$1 fDPk1 gx
O

40

!
Z
gk1 $g3k0 pPk1 0 4

"
qffiffiffiffiffiffiffiffiffiffiffiffiffi
1 $ x2 DPk1 x dx

where the constant Pk1 0 Pk1 x 0 and the function DPk1 x are defined by
Z
Z x %n
Ck0
4 1 1 x2
pffiffiffiffiffiffiffiffiffiffiffiffiffi DPk1 x dx; DPk1 x $
dx
Pk1 0 $
2n1
2
%
3p 0
0 Ok0
1$x

41

42

% k0 corresponds to the dimensionless flux in the zero-viscosity solution (37)


In the above, C
!
Z 1
qffiffiffiffiffiffiffiffiffiffiffiffiffi"
2 %
p1=3
$1
%
%
3cos x x 1 $ x2
Ck0
Ok0 dx xOk0
3
12
x

The explicit analytical expressions for the integrals in (39)(42) can be obtained in the case of a
Newtonian n 1 [21] and a perfectly plastic n 0 fluids, see Appendix D. In the
intermediate case 05n51 the integrals are evaluated numerically.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1452

D. I. GARAGASH

In accordance with earlier discussion of the asymptotic behaviour of the solution near
the crack tip in the general case, the asymptotic behaviour of the small-viscosity solution
(38a)(38c) is given by 37b for the opening and by 23b with Table Ik for the pressure gradient
1 $ x{1 :

1=3

% k p 1 $ x1=2 ;
O
21=2

@Pk
213n=2
$Mk n 1n=3 1 $ x$1n=2
@x
3p

6. TRANSIENT SOLUTION
6.1. Formulation of initial value problem
% k ; Pk ; gk g;
The evolution of the hydraulic fracture solution in toughness scaling, Fk x; Mk fO
with the viscosity parameter Mk from the toughness-dominated regime (Mk 0; zero-viscosity
solution (37)) to the viscosity-dominated regime (Mk 1; zero toughness solution [22]) is
governed by equations in toughness scaling, (16)(24) with Table Ik and 28a :
The initial conditions for this evolution problem are provided by the small-viscosity solution
of the previous section evaluated at a certain initial small (but non-zero) value of viscosity
evolution parameter Mk Mkini ; i.e.
Fk x; Mk Mkini Fk0 x Mkini Fk1 x

43

(Note that the zero-viscosity solution at Mk 0 cannot be used as the starting point of the
transient solution, since the former corresponds to a self-similar solution of the governing
equations, and, therefore, does not allow to resolve the initial Mk -rates of the sought field
variables at Mk 0). Thus, the solution of (16)(24) with Table Ik and 28a ; and initial
conditions (43) is sought for Mk 5Mkini :
6.2. Method of solution
The numerical technique is based on two main steps.
*

The series expansion of the normalized solution for the opening and the net pressure over a
class of base functions in the spatial co-ordinate which satisfy the elasticity singular
integral equation and conform with the proper tip asymptote [15, 22, 33].
Formulation of evolutionary equations (system of ODEs) to solve for the evolution of the
coefficients of the above solution series expansion with Mk :

In the remainder of Section 6.2 we will omit the toughness-scaling index k for brevity.
6.2.1. Spatial series expansion. The solution expansion proposed below is inspired by the
considerations of Adachi [33] and Spence and Sharp [15] for the self-similar solution in the
% and P in the form
Newtonian fluid case n 1: To proceed, we approximate the solution for O
Px; M 4$1 2gM$1=2

N
X

Aj MPnj x AN1 MPnN1 x

44

j0

%
Ox;
M 2gM$1=2 1 $ x2 1=2

N
X

% nj x AN1 MO
% nN1 x
Aj MO

45

j0

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1453

% nj x are selected such that: (i)


where the net pressure and opening base functions Pnj x and O
% nj L$1 fPnj g; (ii) the contribution of
each base function pair satisfies the elasticity equation O
any net pressure base function to the stress-intensity factor is zero
Z 1 n
Pj x dx
pffiffiffiffiffiffiffiffiffiffiffiffiffi 0
46
0
1 $ x2

and (iii) the leading pressure base function Pn0 x provides the net pressure gradient asymptote
at the fracture tip 23b ; i.e. dPn0 x=dx $1 $ x$1n=2 x ! 1 and the corresponding
coefficient A0 M in expansion (44)(45) is given by (24) with Table Ik :
!
"n
2
A0 M Mg1n=2 M GM
47
3
The latter also implies that the net pressure base functions Pnj x with j51 must have either
finite gradient at the tip or a singularity weaker than that of the leading term Pn0 x: In the view
of (i) above, the first constant term in the net pressure expansion (44) yields the expression of the
stress-intensity factor and the corresponding first term in the expression for the opening (45)
satisfies propagation condition 23a with Table Ik :
The leading net pressure base function Pn0 x which satisfies conditions (ii) and (iii) above is
selected as
"
8 1n=2 !
2
>
2 1$n=2
<2
x1 $ x
$
; n=1
2 $ np
1$n
48
Pn0
>
:
2
ln1 $ x 2 ln 2;
n1

% n0 x is obtained from the elasticity integral (22) in the form


The corresponding O
8
0
1
23 $ n
1z
>
3$n
>
z

ln
$
p
cotnpz
>
>
B
C
>
2 $ n
1$z
25n=2
>
B
C
>
<
B
C; n=1
1 A
F
4
$
n;
1;
5
$
n;
$

%On0 1 $ n3 $ np@
2
1
F
$3

n;
1;
$2

n;
z
49
2 1
z
$

>
>
>
3$n
4 $ nz
>
>
>
pffiffiffiffiffiffiffiffiffiffiffiffiffi
>
:
4p $ 8x arcsin x $ 8 1 $ x2 ;
n1
pffiffiffiffiffiffiffiffiffiffiffiffiffi
2
n
% 0 is given by
where z 1 $ x : The near-tip asymptotic behaviour of O
8
np
16 cot 2
>
<
1 $ x3$n=2 O1 $ x3=2 ; n=1
n
1
$
n3
$
n
%
1 $ x{1 : O0
50
>
:
3=2
n1
4p1 $ x O1 $ x ;
According to 23a and (50), the opening asymptotic expansion near the tip contains terms of
order 1 $ x1=2 ; 1 $ x3$n=2 (or 1 $ x for n 1), and 1 $ x3=2 : Therefore, following the
numerical treatment of a fracture driven by a Newtonian fluid in References [15, 33], we select
% nj x with 14j4N in expansion (45) as the weight
the rest of the base opening functions O
2 3=2
function 1 $ x times a polynomial:
% nj 1 $ x2 3=2 Tj$1 1 $ 2x2 ;
O

14j4N

51

where Tj$1 is the Chebyshev polynomial of the first kind of degree j $ 1: After a change of
% nj cos3 y=2 cos j $ 1 y; 14j4N: The net pressure
variable, x sin y=2; (51) becomes O
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1454

D. I. GARAGASH

base functions Pnj ; 14j4N; are obtained from the elasticity integral [33]
8
3
>
>
j1
> cos y;
>
>
8
>
>
<
1
5
Pnj
cos y cos 2y;
j2
>
4
16
>
>
>
>
>
1
3 sinj $ 1y
>
: j $ 11 cos y cos j $ 1y
cos 2y $ 1; 25j4N
4
16 sin y

52

Since dPnj =dx 0; 14j4N; at the fracture inlet x 0; a particular net pressure term
PN1 x 2 $ px with non-zero inlet gradient (and zero contribution to the stress-intensity
factor) is added to expansion (44) to model the non-zero fluid flux at the inlet [22, 33]. The
% nN1 x follows from the elasticity integral (22).
corresponding explicit expression for O
n

6.2.2. Solving for the evolution of the coefficients in the solution series expansion for the generic
case 05n51. The series expansion for the net pressure (44) and opening (45) satisfy elasticity
(22), propagation condition and the net pressure tip asymptote (23) with Table Ik : The solution
for the unknown coefficients in this expansion, Aj M; j51; and the normalized fracture length
gk M ought to be obtained from consideration of lubrication equations (16)(17) and global
fluid balance (19).
Let us rewrite the lubrication equation (16) with Table Ik and 28a as
Z 1
2
% 0 dx F
%
1 $ n
53
O
3
x

% is defined as
where prime 0 @=@ln M and F
!
"
Z 1
%
% dx $ 2 G xO
% $ 1 2G
% C
F
O
3
x

54

Substitution of (45) into (53) reveals that it can be rewritten as


"
#
N
1
X
2
$1=2 0 n
0 n
%
1 $ n 2g
o
Aj o
% x
% j x F
3
j0

55

where the functions o


% n x are defined as
% nj x and o
!
Z 1
Z 1 qffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffiffiffiffiffiffiffiffiffi"
1
2
n
n
n
%
arccos x $ x 1 $ x2
o
1 $ x dx
Oj dx; o
%j
%
2
x
x
The use of the opening expansion (45) in the global fluid balance equation (19) results in an
equation, which can be solved for AN1 :
!
N
X
1
1 $2
$1=2 n
n
g $ 2g
AN1 n
Aj o
56
o
% 0 $
% j 0
o
% N1 0 2
j0

Substitution of (56) into (55) and the use of the expression g0 =g 3=21 $ n+G (which follows
from the definition of G d ln g=d ln t and identity 28a ) yield
N
X
2
# nj x Rx; g; A0 ; . . . ; AN
1 $ n
A0j o
57
3
j0
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1455

where the right-hand side R is

%
&
% g$2 o
# nN1 x 12 2g$1=2 o
# n x G
RF

58

and various functions are defined by


# nN1 x
o

o
% nN1 x
;
o
% nN1 0

# nj x o
# nN1 x;
o
% nj x $ o
% nj 0o

# n x o
# nN1 x
o
% n x $ o
% n 0o

59

Let us elaborate on the stated functional dependence of R in (57). We notice from the original
% (54), (17), and expansion (44)(45)) that R defined by (58) is the
definitions (see definition of F;
known function of x and M-dependent parameters g; G; A0 ; . . . ; AN1 : According to (47), G can
1=n
be expressed as the function of g and A0 ; G A0 M$1=n g$1n=2n $ 2=3: In view of the above
and since AN1 is the known function, (56), of g; A0 ; . . . ; AN ; we obtain functional dependence
of R Rx; g; A0 ; . . . ; AN :
Discretization of the lubrication equation (57) over a set of grid points fxi g; i 1; . . . ; N 1;
in the x 2 0; 1 interval yields N 1 equations
N
X
2
# nj xi Rxi ; g; A0 ; . . . ; AN ;
1 $ n
A0j o
3
j0

i 1; . . . ; N 1

60

# nj 0 0; (59), the corresponding form of (60) if evaluated at the inlet point


Note that since o
x 0 is a non-linear algebraic equation, R0; g; A0 ; . . . ; AN 0: (The latter is simply the
% 0 2g2 $1 :) In order to simplify the task
manifestation of the inlet flux condition (20), Cx
of the numerical solution of (60), we select the grid points fxi g; i 1; . . . ; N 1; inside the
x 2 0; 1 interval (i.e. excluding either the crack inlet or the tip). Then, (60) constitutes a system
of N 1 linear ODEs in terms of the N 1 series coefficients fAj g; j 0; . . . ; N: The
# nj xi g; i 1; . . . ; N 1; j 0; . . . ; N; in
N 1)N 1 matrix of numerical coefficients fo
(60) can be inverted to reduce the linear ODEs (60) to the canonical form with respect to the
derivatives fA0j g; j 0; . . . ; N:
Based on the expressions of g0 =g in terms of G and that of G in terms of g and A0 cited in the
text above, we obtain an additional ODE for the normalized crack length g in the form
"!
#
"1=n
2
A0
2
0
1 $ ng g
$
61
3
3
M g1n=2
The N 2 ODEs (60)(61) and appropriate initial conditions
gMini gk0 Mini gk1 ;

A0 Mini Mini g1n=2 Mini

Aj Mini 0;

j 1; . . . ; N

!
"
2 Mini gk1 n

3 gMini

62

63

are solved simultaneously using the ODE solver in Mathematica software version 4.1 (# 1988
2000 Wolfram Research, Inc.) to obtain the solution gM; A0 M; . . . ; AN M:
Note that initial conditions (62) for g and A0 follow from the small-viscosity solution for g;
(38c) with 37c and (41)(42), while the rest of the initial values of expansion coefficients are set
to zero in (63). Alternatively, we could have obtained Aj Mini j 1; . . . ; N values from the
minimization of the error with which the small-viscosity solution is approximated by the series
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1456

D. I. GARAGASH

expansions (44)(45) at M Mini : For small enough Mini (Mini 10$3 was used in the
calculations), however, the smallness of the estimated values Aj Mini ; j 1; . . . ; N; warrants
the initial state approximation (63).
6.2.3. Solving for the coefficients in the solution series expansion for the special cases (n 0 and
n 1). In the special cases of a perfectly plastic n 0 and Newtonian n 1 fluids, the
lubrication equation assumes non-evolutionary forms (30) and (29), respectively, where M
enters as a parameter only. Consequently, in place of the system of ODEs (60)(61) governing
the solution gM; A0 M; . . . ; AN M for a generic value of fluid index 05n51; we will have
an equivalent system of algebraic equations:
0 Rxi ; g; A0 ; . . . ; AN ;
in place of (60), and equation
A0

i 1; . . . ; N 1

Mg1=2 ;

n0

2
3 Mg;

n1

64

65

in place of (61). In (64), the right-hand side is defined by


8
% dP
O
>
>
n0
>
< $ M dx $ 1;
66
R
>
% 3 dP R 1
>
2 %
O
>
%
:$
$ x O dx $ xO; n 1
3
M dx
%
with P and O given by their series expansions (44) and (45), respectively.
Solution of the system of N 2 non-linear algebraic equations (64)(65) for N 2 unknowns
gM; A0 M; . . . ; AN M at a fixed value of viscosity parameter M is obtained using the
Newton solver in Mathematica software. The solution at the first value of M Mini is obtained
using the initial guess (62)(63) following from the small-viscosity solution. The value of M is
then incremented by a fixed increment DM; and the solution for updated M value is obtained
using the initial guess provided by the solution for the previous M value. The use of a
sufficiently small increment DM then ensures the convergence of the Newton solver.

7. RESULTS AND DISCUSSION


7.1. Dimensionless solution
The numerical transient solution is obtained in the range of the evolution parameter Mk 2
10$3 ; 100+ Mkini 10$3 for fluids with behaviour index n f0; 0:1; 0:2; . . . ; 0:9; 1g (0.1
increment). Results presented in this Section are based on calculations using N 10 terms in
the solution series expansion (44)(45) and a non-uniform spacing between the N 1 grid points
(used in the numerical method to discretize the lubrication equations). The latter is selected to
optimize the convergence of the method and minimize the approximation error over a given
range of Mk : Corresponding details of the numerical error analysis and relevant discussion are
given in Appendix E.
Figure 3 shows the raw data from the numerical transient solution for a hydraulic fracture:
the evolution of the normalized coefficients in the solution series expansion (44)(45) with
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1457

(a)

(b)

(c)

% i Ai =Ai 10$2 ; i 0; . . . ; N 1; for the


Figure 3. Evolution of the normalized series coefficients, A
solution with N 10 regular series terms and for various fluid index: (a) n 0; (b) n 0:5; and (c) n 1:

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1458

D. I. GARAGASH

Table II. Numerical values of coefficients Aj in the


solution series expansion with N 10 base terms for
various fluid behaviour index n 0; 0:5; 1; evaluated at
Mk 0:01:
Mk 0:01

n0

n 0:5

n1

A0 ) 10
A1 ) 103
A2 ) 103
A3 ) 104
A4 ) 105
A5 ) 105
A6 ) 105
A7 ) 106
A8 ) 106
A9 ) 106
A10 ) 107
AN1 ) 103

9.5830
$8.8862
2.0589
$1.7606
7.5824
$2.0941
1.0526
$3.4563
1.6632
$0.4262
1.5684
8.4945

7.6040
$8.4974
2.1922
$3.7689
13.814
$5.1754
2.2812
$9.2360
3.7709
$1.2108
3.2491
12.3

12.1
$7.1535
0.1679
$2.9273
9.5975
$5.6495
2.3839
$12.166
4.5690
$1.7804
3.8359
4.0856

viscosity parameter Mk in loglog scale for three values of the fluid behaviour index (a) n 0; (b)
n 0:5; and (c) n 1: The plots of coefficients shown in the figure are normalized by their respective
% j Mk Aj Mk =Aj 0:01: The corresponding values Aj 0:01 are
values at Mk 0:01; i.e. A
given in the Table II. (Thus, the data in the table and in Figure 3 is sufficient to reconstruct the
% Mk ; Px; Mk ; gk Mk g with the help of (44)(45) and (56)). Interestingly,
complete solution fOx;
% j Mk Mk j 0; . . . ; N 1 for Mk 90:1; see Figure 3(a)(b).
for n 0 and n 0:5; A
Consequently, the solution is linear in Mk and, therefore, given by the small viscosity solution (38a)
(38c). For n 1 (Newtonian fluid), the coefficients corresponding to the leading terms in (44)(45)
%0A
%1A
% N1 Mk for Mk 90:01: This
are approximately linear for small Mk ; namely A
indicates applicability of the small-viscosity solution in the case of Newtonian fluid only for small
Mk : (For this case, Reference [21] provides an improved small-viscosity solution with dramatically
increased Mk -range of applicability.)
Profiles of the dimensionless net pressure and opening along the crack for gradually
increasing values of viscosity parameter Mk from 0:01 to 10 are shown on Figures 4 and 5,
respectively, for three different values of fluid index (aa0 ) n 0; (bb0 ) n 0:5; and (cc0 ) n 1:
% k ; (ac), and viscosity-scaling, Pm
Solution is shown in both toughness scaling, Pk and Ok gk O
% m ; (a0 c0 ). The transient solution in toughness scaling, plots (ac) in Figures 4 and
and Om gm O
5, and in viscosity scaling, plots (a0 c0 ) in Figures 4 and 5, converges to the small-time asymptote
(zero-viscosity solution, (37)) and to the large-time asymptote (zero-toughness solution [22]) in
the limit of small and large viscosity parameter Mk ; respectively. (The above asymptotes are
shown by dashed lines.) The corresponding ranges of viscosity where the transient solution is
adequately approximated by its small and large-time asymptotes can be estimated by Mk 40:01
(toughness-dominated range) and Mk 510 (viscosity-dominated range). Comparison of
solutions for different fluid index indicates that the above toughness-dominated range shrinks
and the above viscosity-dominated range expands with the increase of the fluid behaviour index.
These dependence of the extent of asymptotic regions of the transient solution on fluid
index n is also evident on Figure 6 which shows the evolution of the overall characteristics
of the transient fracture solution, namely, (aa0 ) dimensionless fracture half-length, (bb0 )
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

(a)

(a)

(b)

(b)

(c)

(c)

1459

Figure 4. Profiles of dimensionless net-pressure Pk;m along the fracture for fluid behaviour index
(aa0 ) n 0 (plastic fluid), (bb0 ) n 0:5; and (cc0 ) n 1 (Newtonian fluid) and various values of
the evolution parameter Mk f0:01; 0:0316; 0:1; 0:316; 1; 3:16; 10g: Figures (ac) and (a0 c0 )
correspond to the toughness and viscosity scalings, respectively. Dashed lines show (ac) zeroviscosity solution and (a0 c0 ) zero-toughness solution.

dimensionless net pressure at the inlet and at the tip, and (cc0 ) dimensionless crack opening at
the inlet, for various values of the fluid behaviour index from n 0 to n 1 (in 0:1 increments).
Plots in Figure 6(ac) show the evolution of respective field variables in toughness-scaling from
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1460

D. I. GARAGASH

(a)

(a)

(b)

(b)

(c)

(c)

Figure 5. Profiles of dimensionless opening Ok;m along the fracture for the set of parameters of Figure 4.

the early time asymptotic constants corresponding to the zero-viscosity solution and plots on
Figure 6(a0 c0 ) show the evolution of respective quantities in viscosity-scaling towards the
asymptotic constants corresponding to the zero-toughness solution [22]. The solution departs
from the toughness-dominated regime, see Figure 6(ac), and approaches the viscositydominated regime, see Figure 6(a0 c0 ), faster for higher values of fluid index.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

(a)

(a)

(b)

(b)

(c)

(c)

1461

Figure 6. Evolution of the solution in toughness scaling, (ac), and viscosity scaling, (a0 c0 ), with the
dimensionless viscosity Mk for different values of fluid behaviour index n 0; . . . ; 1 (in increments of 0:1):
a; a0 dimensionless half-fracture length gk;m ; b; b0 ); dimensionless net pressure at the inlet Pk;m 0 and
% k;m 0:
fracture tip Pk;m 1; c; c0 ), dimensionless opening at the inlet O

Profiles of the normalized net pressure on Figure 4 confirm that for shear-thinning fluids
(n51) the net pressure is finite at the fracture tip (see also direct plot of the normalized tip net
pressure on Figure 6(bb0 )) with its value in the toughness-scaling decreasing from Pk0 p1=3 =8
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1462

D. I. GARAGASH

at zero time Mk 0 to eventually negative infinity at infinite time (Mk 1 or Km 0). The
unbounded large-time limit of the normalized net pressure at the tip in the viscosity scaling
(which is the appropriate scaling for the large-time viscosity-dominated fracture propagation) is
an artefact of different time-dependence of the net pressure at the crack inlet and at the tip,
respectively. Indeed, in the large-time limit, the dimensionless net pressure at the inlet converges to a
finite positive O1 value given by the zero-toughness solution (Figure 6b0 , and, therefore, the timedependence of the dimensional net pressure at the inlet is prescribed by the viscosity scaling, (13b),
t!1:

p0 em tE 0 Pm 0 ' t$n=2n

(see Equation 14a and Table CI for Lm t). On the other hand, analysis of the LEFM tip
boundary layer scaling in Appendix B for the case of a shear-thinning fluid suggests that at the tip
t!1:

# # ' $t$n=4$n2
p e#tE 0 P
jx0

67

(see Equations (B9), (B7)b with vtip ' L m and Table CI for Lm t). Consequently, the rate
of decay of the inlet net pressure with time exceeds that of the tip net pressure, resulting in
p=p0 ! 1 as t ! 1:
The net pressure at the crack tip decreases from its early-time maximum positive value in the
toughness-dominated regime to intermediate negative values in the transient part of the solution
(as in Figure 6(bb0 ) for the normalized pressure). However, this trend has to be eventually
reversed, as the dimensional solution approaches the large-time viscosity-dominated asymptote
(67), where the absolute value of the negative net pressure at the tip vanishes with time. This
behaviour suggest the existence of the global minimum of the (negative) dimensional net pressure
at the tip in the transient solution, and rather interesting possibilities for the occurrence and
evolution of the fluid lag. Indeed, a fluid lag can develop at some point of the fracture evolution if
the fluid pressure at the tip drops below the cavitation value (e.g. when the absolute value of the
negative net pressure at the tip exceeds the confining stress, $ p > so ; see discussion in Section
3), and then to increase as the fraction of the fracture length for at least some portion of the
fracture evolution. On the other hand, the fluid lag has to be exactly zero in the limit of infinite
time (infinite fracture length), since p ! 0: Therefore, the transient fluid lag has to evolve nonmonotonically, increasing from zero at the onset of the tip cavitation to a certain maximum
fraction of the fracture length and then decreasing to zero. This type of behaviour is qualitatively
different from the behaviour of the fluid lag in hydraulic fracture driven by a Newtonian fluid
where the lag monotonically decreases with time from the early time maximum value [30].
7.2. Example of a dimensional solution for the crack length and the net pressure evolution
Figure 7 illustrates the evolution of the dimensional fracture half-length with time t in loglog
scale for hydraulic fractures driven by fluids with various behaviour index n (between 0 and 1 in
increments of 0.2). Plots on Figure 7 were obtained from the plot of dimensionless half-length gk
on Figure 6(aa0 ) by appropriate rescaling using the following set of problem parameters
representative of reservoir hydraulic fracturing applications: E 25 GPa; n 0:15; Qo
10$3 m2 =s; and three different values of material toughness: (a) KIc 1 MPa m1=2 ; (b) KIc
2 MPa m1=2 ; and (c) KIc 5 MPa m1=2 ; respectively. The values of consistency index M are
selected different for fluids with different behaviour index n; such that the rheological curves for
the considered set of fluids intersect at a common (apparent) value of viscosity mn sxy =g at a
given reference value of shearing rate g [22]. The apparent viscosity of these fluids is set for this
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1463

(a)

(b)

(c)

Figure 7. Evolution of the crack half-length with time for different values of fluid behaviour index from
n 0 (plastic fluid) to n 1 (Newtonian fluid) in 0:2 increments and three different values of toughness
KIc 1; 2; and 5 MPa m1=2 : Open incremental symbols indicate 5% departure from the corresponding zerotoughness solution [22]. Zero-toughness solution for n 0 is shown by the short-dash line. Zero-viscosity
solution (independent of n) is shown by the long-dash line.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1464

D. I. GARAGASH

example to be equal to mn 0:1 Pa s at a reference value of shearing rate g 50 s$1 :


Consistency index M of each fluid is therefore calculated via M mn g 1$n :
Figure 7 shows the departure from the early-time toughness-dominated regime (zero-viscosity
solution shown by the long-dash line) and convergence towards the large-time viscositydominated regime (zero-toughness solution [22] shown for n 0 by the short-dash line). Open
incremental symbols on Figure 7(a) and (b) indicate the time at which the transient solution is 5%
away from the zero-toughness large-time limit. Importantly, we observe that for considered set of
problem parameters and moderate toughness values 122 MPa m1=2 ; fracture length evolution is
effectively given by the zero-toughness asymptote in the practical time (fracture length) range for
fluids with high enough index n: For example, the solution is within 5% of the zero-toughness
asymptote in the whole range of time/length represented in Figure 7(a) for n 0:8 and 1
(Newtonian fluid) and in Figure 7(b) for n 1: However, the transient solution is significantly
different from its large-time asymptote for fluids with low enough index n: For example, the
transient solution for KIc 1 MPa m1=2 (Figure 7(a)) is within 5% of the zero-toughness
asymptote for t0135 s > 50 m for n 0 and for t030 s > 20 m for n 0:2: For a two-fold
increase of material toughness to the value KIc 2 MPa m1=2 (Figure 7(b)), the transient solution
is significantly different from its large-time asymptote for cracks as long as 200 m for fluids
with index n 0 and 2 m for fluids with index as large as n 0:8: Finally, for values of
toughness as large as KIc 5 MPa m1=2 (Figure 7(c)), the transient solution is significantly
different from its large-time asymptote in the entire shown time range for all fluids of interest.
Latter observations drawn from Figure 7 emphasize the relevance of the transient solution in
the practical treatment time (fracture length) range for shear-thinning fluids with index n90:6
for KIc 122 MPa m1=2 : For higher apparent values of toughness, which are often considered
relevant for large-scale fracture propagation under typical level of in situ confining stress (tens of
MPa) [34, 35] and/or in damaged rock [36, 37], the transient solution becomes evermore
important for the entire spectrum of shear-thinning fluids n51:
Figure 8 illustrates the evolution of the dimensional net pressure at the inlet p0 with the crack
half-length in the loglog scale for the hydraulic fracture propagation example considered
above with KIc 2 MPa m1=2 : This plot can be conceivably used for the inversion of the
bottom hole net pressure, p0; to estimate the fracture length. At early time, fracture
propagates in the toughness-dominated regime and decline of the inlet net pressure p0 follows
from the toughness scaling: p0 ' $1=2 ( ' t2=3 ; p0 ' t$1=3 ), see (13a)(13e) and Table Ik :
At large time, propagation in the viscosity-dominated regime yields the inlet pressure decline as
p0 ' $n=n1 ( ' tn1=n2 ; p0 ' t$n=n2 ), see (13a)(13e) and Table Im : Consequently,
for all shear-thinning fluids n51 the pressure decline decreases with fracture growth n=n 1
51=2; see Figure 8. In the perfectly plastic fluid case n 0 the inlet net pressure levels off at
a fraction of MPa for fracture half-length exceeding 100 m:
Finally, we consider the evolution of the dimensional net pressure at the crack tip p with the
crack half-length in the linear-log scale (Figure 9) for the example considered above with
KIc 2 MPa m1=2 and fluid index in the range between n 0 and n 0:8 (recall that
p $1; for a Newtonian fluid, n 1). The figure illustrates the premise discussed earlier in
this section that the net pressure at the tip evolves in the non-monotonic fashion: initially falling
from positive to negative values, passing a global (negative) minimum and asymptotically
increasing to zero value in the large-time limit. In the shown range of the crack length, the global
net pressure minimum can be observed for fluids with n 0:4 and 0.5. (The minimum of the tip
net pressure takes place later or earlier in the crack length evolution than shown on the figure for
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1465

Figure 8. Net-pressure at the crack inlet p0 vs the half-length for the set of parameters of
Figure 7 and KIc 2 MPa m1=2 :

Figure 9. Net pressure at the crack tip p vs the half-length for KIc 2 MPa m1=2 and
various values of the fluid index n:

fluids with the index below n 0:4 or above n 0:5; respectively.) Depending on the cavitation
threshold $so for the net pressure, the fluid lag can emerge during intermediate stages of
fracture evolution and disappear as the fracture approaches the large time, zero tip net pressure
limit. For the considered hydraulic fracture example (Figure 9), the minimum tip net pressure
for fluids with n40:5 exceeds $2 MPa level and, therefore, precludes cavitation with exception
for shallower depths (with so 52 MPa). Figure 9 also shows that for larger fluid index n
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1466

D. I. GARAGASH

(approaching the limit of Newtonian fluid n 1), the cavitation and corresponding fluid lagging
behind the fracture tip may become eminent at early stages of fracture propagation (depending
on the level of so ). To arrive to more general conclusions about the relevance of the fluid lag for
the fracture propagation driven by a non-Newtonian shear-thinning fluid would require a future
analysis explicitly accounting for the fluid lag at the fracture tip (akin to that in References
[30, 38] devoted to the Newtonian fluid case).
8. CONCLUSIONS
A transient solution for the problem of a fracture driven by the injection of a power-law fluid in
an impermeable, linear elastic rock of finite toughness has been constructed. The numerical
method combines the technique of the solution series expansion in space, developed earlier
[15, 22] in applications to self-similar hydraulic fracture propagation problems, with the explicit
ODEs formulation for the evolution of the series expansion coefficients in time. The analysis of
the problem scaling and of the numerical transient solution has shown that the fracture evolves
from the early-time toughness-dominated regime to the large-time viscosity-dominated regime.
The fracture propagation in the above two limiting regimes can be associated with the
propagation of a fracture driven by the respective equivalent Newtonian fluids characterized
by time-dependent viscosities. The approximate ranges of material parameters and time (lumped
into a single non-dimensional evolution parameter Mk m0k tE 03 Qo =K 04 with the characteristic
1=4
time-dependent viscosity m0k t M 0 K 0 =E 0 Qo 41$n=3 t21$n=3 corresponding to these two
distinct asymptotic regimes were discussed. Particularly, it has been shown that a decrease of
fluid behaviour index n (which corresponds to the increasing shear-thinning property of the
fluid) results in the increase of the time/parametric domain for toughness-dominated regime and
decrease of corresponding domain for the viscosity-dominated regime. In the particular example
of problem parameters relevant to hydraulic fracturing treatments in the petroleum industry, it
was shown that hydraulic fracture propagation occurs in the transient regime (i.e. significantly
different from either two limiting regimes) for almost the entire time-range (fracture lengthrange) of practical interest for small enough index n (n90:4 for the considered set of problem
parameters and moderate value of rock toughness KIc 2 MPa m1=2 ). Whereas, for large
enough index n n00:8 hydraulic fracture propagation occurs in the viscosity-dominated
regime in the time-range of interest.
APPENDIX A: ACCOUNT OF THE LUBRICATION THEORY
FOR POWER-LAW FLUIDS
Under the lubrication assumption [26], the fluid inertia and the variation of the shear stress in
the flow direction (along the x-axis in Figure 1) are neglected, compared to the shear stress
variation across the flow channel (along the z-axis perpendicular to the fracture plane) and to
the driving pressure gradient, resulting in the equations of fluid motion in the form
@p @sxz
@p
0$
A1
; 0$
@x
@z
@z
Using the symmetries with respect to the x- and z-axis, let us consider the positive co-ordinate
quadrant (x50; z50) where the absolute values of the shear stress and the shear-strain rate are
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1467

t $sxz and g $2exz $@vx =@z; respectively. (Contribution ($@vz =@x) to the shear strainrate g is neglected in the above expression under the lubrication assumption.) Upon substitution
of the above expressions and of the constitutive law (1) in A1a ; integrating once in z; and using
the symmetry condition g z 0 0; one finds
@vx

g $
@z

!
"
z @p 1=n
$
M @x

A2

where gradient @p=@x is the function of the channel co-ordinate x (and time) only, A1b :
Poiseuilles law (3) [23] for the fluid flux in the channel can be obtained from the identity
q2

w=2

vx dz $2

w=2
0

! "
@vx
z dz
@z

A3

(based on the channel symmetry, integration by parts, and the stick condition along the channel
wall, vx z w=2 0) and expression for the shear rate (A2).
Let us now evaluate the apparent fluid viscosity mn t=g M g n$1 (see (1)) at the
fracture channel wall z w=2: Using the expression for the pressure gradient j@p=@xj
M 0 =wjqj=w2 n following from Poiseuilles law (3) (where M 0 fM) in the expression for
g z w=2 following from (A2), and substituting the result into the expression for the apparent
viscosity, we find

mwall

! 2 "1$n
n
0 w
M

42n 1
jqj

A4

Consequently, the Poiseuilles law for a power-law non-Newtonian fluid, (3), can be expressed in
the form of the Poiseuilles law for an equivalent Newtonian fluid, (4), with an effective
viscosity m0 4n$1 2n 1mnwall :

APPENDIX B: LEFM BOUNDARY LAYER CONSIDERATIONS


In this Appendix we present the scaling considerations for the LEFM boundary layer, which
exists at the tip of an arbitrary finite fluid-driven fracture under small toughness conditions.
These considerations are direct extension of the analogous considerations for the case of
#
Newtonian fluid [19]. The solution in the boundary layer of thickness {
can be approximated
to the leading order by the solution of a semi-infinite crack propagating at the velocity vtip given
by the instantaneous crack tip velocity in the underlined problem of a finite crack, vtip
(Figure 2). Governing equations for a semi-infinite steadily propagating crack with zero fluid lag
(e.g. Reference [10]) are presented here for an equivalent Newtonian fluid with effective
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1468

D. I. GARAGASH

viscosity m0 in the form


# dp
w3 x
#
qx;
# dx#
m0 x
1
# E
px
4p
0

# M
m x
Z

#
w2 x
#
qx

dws ds
;
ds x# $ s

"1$n

# wxv
# tip
qx

x# ! 0 : w

K 0 1=2
x#
E0

B1

B2

where x# xtip $ x is a co-ordinate moving with the crack tip (Figure 2). Equations (B1)a2c
specialize the fluid flux q; effective viscosity m0 ; and continuity under lubrication flow
approximation, respectively; Equations (B2)a2b specialize elasticity integral relation between the
net pressure and the crack opening and the LEFM crack tip asymptote, respectively. The
solution of above equations for a semi-infinite steadily propagating crack provides the boundary
layer solution which governs the transition from the LEFM near-tip asymptote (w ' x# 1=2 for
# to the viscosity-dominated away from the tip asymptote (w ' x# 2=2n for xc
# see the
# )
# ),
x{
structure of the boundary layer solution on Figure 2.
The tip boundary layer scaling can be defined in terms of the characteristic tip lengthscale #
(boundary layer thickness), characteristic crack near-tip aspect ratio e#; and characteristic tip
viscosity m# 0 as
# x
#
# wx
# x;
#
#
# x;
# e## O
# e#E 0 P
B3
x# # x;
px
m0 m# 0 M
# are dimensionless
# P;
# and M
where x# is the normalized distance from the crack tip, and O;
opening, net-pressure, and effective viscosity in the boundary layer scaling, respectively. The
choice of scaling parameters # and e# for an equivalent Newtonian fluid with characteristic
viscosity m# 0 (see Equation (35) of Reference [19])z
0
16
0
0 0
K
A ; e# E m# vtip
# @
B4
1=3
K 02
E 02=3 m# 01=3 v
tip

and the following implicit expression for m# 0 :

! # "1$n
e#
m# M
vtip
0

reduce boundary layer equations (B1)(B2) to the following normalized form:


Z 1 #
#
dOs ds
# x# 1=2
# 1n dP 1; P
# 1
# x
; x# ! 0 : O
O
4p 0
ds x# $ s
dx#

B5

B6

# O
# 1$n has been used to arrive to (B6)a ).
(Expression for the normalized effective viscosity M
Solution of normalized equations (B6) as well as the elaborate studies of the near tip and
away from the tip asymptotic expansions of the solution have been recently carried out by
Garagash et al. [14] for the case of Newtonian fluid n 1 and by Kresse et al. [39] for the case
of a general power-law shear-thinning fluid (05n51). Remarkably, Equations (B6) are
parameter-free for a given fluid index n; and, therefore, the scaling (B3) with (B4)(B5)
completely defines the dependence of the tip boundary layer solution on material parameters
z

Note the corrected typo in the expression for e# in the 2nd of Equation (35) of Reference [19].

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1469

and time (via the time-dependence of the crack tip velocity vtip of the underlined finite fracture
problem). It is, therefore, useful to expand the implicit definitions (B4)(B5) of characteristic
# characteristic crack tip aspect ratio e#; and characteristic tip viscosity
boundary layer thickness ;
m# 0 to their explicit form
!2=2$n
!1$n=2$n
!
"1=2$n
02n
04
M 0 vntip
K
K
; e#
; m# 0 M 01=2$n
#
E 01n M 0 vntip
K 02n E 01$2n
E 03 v2tip
B7
The LEFM tip boundary layer exists near the front of a finite fracture when the extent of the
boundary layer is much smaller than the global fracture dimension (e.g. half-length of a planar
#
fracture), {:
Similar to Reference [19], we can define the dimensionless tip toughness
parameter as
! # "2$n=22n
K0
#m
K

B8
n=2n 2$n=22n

E 01n=2n M 01=2n vtip

Consequently, the condition when the LEFM boundary layer exists near the tip of a finite
fracture can be formulated as a condition on smallness of the dimensionless tip toughness,
22n=2$n
#m
# K
=
{1: Under this condition, the solution away from the tip boundary layer
#
# is approximately independent of toughness and is given by the zero-toughness outer
xc
solution (see the sketch on Figure 2).
The tip boundary layer scaling and corresponding solution of normalized governing
equations (B6) can yield useful information about the tip behaviour of a finite fracture
propagating in viscosity-dominated regime (i.e. under small toughness conditions). For
example, Equations (B6) imply that the normalized net pressure is finite and negative at the
# x# 0 O1 given by
crack tip for shear-thinning fluids, with the numerical value P
the numerical solution in Reference [39]. Thus, the time evolution of dimensional net pressure at
the tip of a finite fracture propagating in the viscosity-dominated regime (under small-toughness
conditions) is given by
# #
ptip e# tE 0 P
jx0

B9

In view of (B7)b ; the evolution of the net pressure at the crack tip dominated by the LEFM
boundary layer solution is defined by that of the crack tip velocity vtip ; and, specifically, vanishes
in the limit of slowing fracture, vtip ! 0; typical of fractures driven by a constant fluid injection
rate.

APPENDIX C: SCALING EXPRESSIONS EXPANDED IN TIME


The toughness and the viscosity scalings (i.e. corresponding choices of scaling parameters e and
L; and dimensionless groups M and K), presented in Table I as the scalings for the equivalent
Newtonian fluids with time-dependent viscosities m0k t and m0m t; respectively, are expanded in
Table CI to reveal explicit time-dependence of the scaling parameters and dimensionless groups.
Corresponding expressions for the scaling parameter e follow from 14a and Table CI.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1470

D. I. GARAGASH

Table CI. Quantities corresponding to the viscosity (M) and the toughness (K) scalings.
Scaling
K
M

L
!

"2=3

1=4

E 0 Qo t
M 0 E 0 Qo
0
E0
K0
K
! 0 "1=2n2
1=2 E
tn1=n2
Qo
M0

M
!4=32n

K
t2=31$n

1
K0

1=4

E 0 Qo

E0
M0

"3=42n

t$1=21$n=2n

APPENDIX D: EXPLICIT FORMULAE IN THE SMALL VISCOSITY SOLUTION


% 1 ; P1 ; g1 g; (39)(42), in the small-viscosity solution (38a)(38c)
The first-order viscosity term fO
can be evaluated explicitly in the case of a perfectly plastic (n 0) and a Newtonian (n 1; see
Reference [21]) fluids. In the former case, integration in (39)(42) yields
n0:

Pk1

4 3p2
2
$ 1=3 arcsin x
4=3
p
6p

% k1 8 6p
O
3p4=3

qffiffiffiffiffiffiffiffiffiffiffiffiffi
8
1 $ x2 $ 4=3 Isin x
p

gk1 $
where the function Isin x is defined by
Z 1
Z
Isin x
Gx; x0 sin$1 x0 dx0

64
3p7=3

D1

D2
D3

$
$
$cos a cos a0 $ 0
$a cos a0 da0
ln$$
cos a $ cos a0 $
0
0
! 2
"
a
p
$ 2 ln tan a
2 ln tan cos a
2
4
p=2

sin a ImLi2 e2ia $ Li2 $e2ia ;

a arcsin x

D4

where Li2 is the dilogarithmic function. The value of the opening at the crack inlet follows
% k1 0 81 ln 64=3p4=3 :
from (D2) and (D4), O
For a Newtonian fluid, the first-order term is reproduced below after Reference [21] for
completeness
!
qffiffiffiffiffiffiffiffiffiffiffiffiffi
8
1
3 x arccos x
2
ln4 1 $ x $ pffiffiffiffiffiffiffiffiffiffiffiffiffi
n 1 : P1 2=3
D5
24
4 1 $ x2
3p
2
pffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffi2ffi 31
!
"qffiffiffiffiffiffiffiffiffiffiffiffiffi
3 41 1 $ x2 1 1$x 5A
2
% 1 8 @2p $ 4x arcsin x $ 5 $ ln 2
ln
1
$
x
$
O
pffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffi2ffi
6
2
3p2=3
1 $ 1 $ x2 1$ 1$x
0

g1 $
Copyright # 2006 John Wiley & Sons, Ltd.

321 6 ln 2
$2:7220
9p5=3

D6

D7

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1471

APPENDIX E: NUMERICAL ERROR ANALYSIS


The numerical calculations of the transient solution are performed for various number of terms
in the series expansion (44)(45), 54N410 to investigate the convergence of the numerical
method and the approximation error in the lubrication equation. The spacing between the
N 1 grid points along the crack f05xi 51g; i 1; . . . ; N 1; used in the numerical method
to discretize the lubrication equations (resulting in (60) for 05n51 and (61) for n 0 or n 1),
is selected to optimize the convergence of the method and minimize the approximation error
over a given range of Mk 2 10$3 ; 100+:
For a given solution series expansion (44)(45) with N base terms, we define the local
error and global quadratic error in the lubrication equation (16) with Table Ik and
28a by
Z 1
RHSN x; Mk
N
EN x; Mk 1 $
M

EN x; Mk +2 dx
E1
;
e
k
LHSN x; Mk
0
respectively, where RHSN x; Mk and LHSN x; Mk are the right- and left-hand sides of the
lubrication equation (16) evaluated at the given solution series expansion.
For each 54N410 we have considered different sets of N 1 grid points parametrized by a
parameter s 2 0; 1+ as follows:
8 $3
10 ;
i1
>
<
'
(
'
!
"
(
2
E2
xi
i$1
i$1
>
:s
; i 2; . . . ; N 1
1 $ s 1 $ 1 $
N 1
N1

Figure E1. Distributions of lubrication equation error E10 x; Mk along the crack at Mk 1
corresponding to the numerical solutions with different sets of grid points (see the insert) with grid
parameter s 0:1; s 0:3; and s 0:5; respectively.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1472

D. I. GARAGASH

(a)

(b)

Figure E2. Quadratic global error eN as a function of evolution parameter Mk : (a) for the fixed value
n 0:5 of fluid index and various number of regular series terms N; (b) for fixed number of terms N 10
and various values of fluid index n from 0 (plastic fluid) to 1 (Newtonian fluid).

Grid points xi i 2; . . . ; N 1 are regularly spaced for s 1 and concentrated toward the
fracture tip xN2 1 for s51; see the insert on Figure E1 showing sets (E2) for N 10 and
various s: The parameter s controls the concentration of grid points near the tip. Example of
the spatial distributions of the local lubrication error EN x; Mk 1 for calculations with
N 10 for three different sets of grid points with s 0:5 (small concentration), s 0:3; and
s 0:1 (large concentration) is shown on Figure E1. (Note that the solution with uniformly
spaced grid points, s 1; does not converge, and, thus, is not shown.) Corresponding global
quadratic error e10 Mk 1 values are 1:4 ) 10$5 ; 9:7 ) 10$6 ; and 3:8 ) 10$6 ; respectively.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1473

Clearly, the piled-up set with s 0:1 corresponds to the lowest global lubrication error, and to
the mostly uniform local error distribution along the crack. (Large spikes in the local error is
observed near the tip of the crack for the solutions with s 0:3 and s 0:5:) Other numerical
tests for various fluid behaviour index n and various values of Mk have shown the same trends
of the error dependence on the grid parameter s: Consequently, the grid with s 0:1 has been
used for all of the subsequent calculations.
Figure E2 shows the variation of the global lubrication error eN with Mk for (a) various
number of terms N f5; 6; 7; 8; 9; 10g in the solution series expansion for a fixed fluid index
n 0:5; (b) various values of the fluid index n f0; 0:1; . . . ; 0:9; 1g (0.1 increment) and the
fixed number N 10 of terms in the series. From Figure E2(a) we observe that the lubrication
error naturally decreases with the number of terms in the series, and interestingly, the
significant error reduction occurs when increasing N from odd to even number (e.g. from
N 7 to N 8) and not from even to odd (e.g. from N 8 to N 9). Figure E2(b) indicates
a general trend of reduction in numerical accuracy of the solution with increasing fluid index n;
with the most accurate solutions obtained for the case of perfectly plastic fluid n 0: With
respect to the dependence of the solution accuracy on the normalized viscosity Mk ; we observe
that the error increases with Mk and the numerical solution (especially for higher values of
fluid index) cannot be considered substantially accurate for Mk larger than 10 (e.g. for n 1
and Mk > 10 the quadratic error exceeds 10%). This loss of accuracy of solution in
approaching the zero toughness (infinite viscosity) limit is related to the diminishing influence
of toughness at the fracture tip and the emerging viscosity near-tip asymptote, characterized
% ' 1 $ x2=2n and the net pressure singularity P ' 1 $ x$n=2n at the tip,
by opening O
1 $ x{1; (see (12) or (33) in viscosity scaling). The numerical method relying on the toughness
asymptote at the tip breaks down in the large viscosity limit. Similar observation has been
made in References [15, 33] for the case of a Newtonian fluid. In order to accurately capture
the solution for large viscosity, one need to explicitly consider the near-tip boundary layer
(Section 3 and Appendix B) which provides the transition from the toughness-dominated
behaviour immediately next to the tip to the viscosity-dominated behaviour at intermediate
distances from the tip [19].
ACKNOWLEDGEMENTS

Acknowledgement is made to the Donors of The Petroleum Research Fund, administered by the American
Chemical Society, for partial support of this research under Grant ACS-PRF 36729-G2. The author also
wishes to acknowledge Jose I. Adachi, Emmanuel Detournay, and Leonid Germanovich for their helpful
review comments which led to considerable improvement in the paper.

REFERENCES
1. Economides MJ, Nolte KG. Reservoir Stimulation (3rd edn). Wiley: Chichester, U.K., 2000.
2. Spence DA, Turcotte DL. Magma-driven propagation crack. Journal of Geophysical Research 1985; 90:
575580.
3. Rubin AM. Propagation of magma-filled cracks. Annual Review of Earth and Planetary Sciences 1995; 23:
287336.
4. Kristianovich SA, Zheltov YP. Formation of vertical fractures by means of highly viscous fluids. Proceedings of the
4th World Petroleum Congress, vol. 2, Rome, 1955; 579586.
5. Barenblatt GI. On the formation of horizontal cracks in hydraulic fracture of an oil-bearing stratum. Prikladnaya
Matematika i Mekhanika 1956; 20:475486.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

1474

D. I. GARAGASH

6. Perkins TK, Kern LR. Widths of hydraulic fractures. Journal of Petroleum Technology 1961; 222:937949.
7. Geertsma J, De Klerk F. A rapid method of predicting width and extent of hydraulic induced fractures. Journal of
Petroleum Technology 1969; 246:15711581.
8. Nordgren RP. Propagation of vertical hydraulic fractures. Journal of Petroleum Technology 1972; 253:306314.
9. Abe H, Mura T, Keer LM. Growth rate of a penny-shaped crack in hydraulic fracturing of rocks. Journal of
Geophysical Research 1976; 81:53355340.
10. Desroches J, Detournay E, Lenoach B, Papanastasiou P, Pearson JRA, Thiercelin M, Cheng AHD. The crack tip
region in hydraulic fracturing. Proceedings of the Royal Society of London, Series A 1994; 447:3948.
11. Lenoach B. The crack tip solution for hydraulic fracturing in a permeable solid. Journal of the Mechanics and Physics
of Solids 1995; 43:10251043.
12. Garagash D, Detournay E. The tip region of a fluid-driven fracture in an elastic medium. Journal of Applied
Mechanics 2000; 67:183192.
13. Detournay E. Propagation regimes of fluid-driven fractures in impermeable rocks. International Journal of
Geomechanics 2004; 4:111.
14. Garagash DI, Detournay E, Adachi JI. Tip solution of a fluid-driven fracture in permeable rock. Journal of the
Mechanics and Physics of Solids 2006, submitted.
15. Spence DA, Sharp P. Self-similar solutions for elastohydrodynamic cavity flow. Proceedings of the Royal Society of
London, Series A 1985; 400:289313.
16. Lister JR. Buoyancy-driven fluid fracture: the effects of material toughness and of highly viscous fluids. Journal of
Fluid Mechanics 1990; 210:263280.
17. Carbonell RS, Desroches J, Detournay E. A comparison between a semi-analytical and a numerical solution of a
two-dimensional hydraulic fracture. International Journal of Solids and Structures 1999; 36:48694888.
18. Savitski AA, Detournay E. Propagation of a fluid-driven penny-shaped fracture in an impermeable rock: asymptotic
solutions. International Journal of Solids and Structures 2002; 39:63116337.
19. Garagash DI, Detournay E. Plane strain propagation of a fluid-driven fracture: small toughness solution. Journal of
Applied Mechanics 2005; 72:916928.
20. Garagash DI. Hydraulic fracture propagation in elastic rock with large toughness. In Rock Around the
Rim}Proceedings of the 4th North American Rock Mechanics, Girard J, Liebman M, Breeds C, Doe T (eds).
Balkema: Rotterdam, 2000; 221228.
21. Garagash DI. Plane strain propagation of a fluid-driven fracture during injection and shut-in: asymptotics of large
toughness. Engineering Fracture Mechanics 2006; 73:456481.
22. Adachi JI, Detournay E. Self-similar solution of a plane-strain fracture driven by a power-law fluid. International
Journal for Numerical and Analytical Methods in Geomechanics 2002; 26:579604.
23. Ben-Naceur K. Modeling of hydraulic fractures. In Reservoir Stimulation (2nd edn), Economides MJ, Nolte KG
(eds). Prentice-Hall: Englewood Cliffs, 1989.
24. Cameron JR, Prudhomme RK. Fracturing-fluid flow behaviour. In Recent Advances in Hydraulic Fracturing,
Gidley JL, Holditch SA, Nierode DE, Veatch RW (eds). Society of Petroleum Engineers: Richardson, 1989;
177209.
25. Webb SL, Dingwell DB. Non-Newtonian rheology of igneous melts at high stresses and strain rates:
experimental results for rhyolite, andesite, basalt, and nephelinite. Journal of Geophysical Research 1990;
95(10):15,69515,701.
26. Batchelor GK. An Introduction to Fluid Dynamics. Cambridge University Press: Cambridge U.K., 1967.
27. Bilby BA, Eshelby JD. Dislocations and the theory of fracture. In Fracture, an Advanced Treatise, Liebowitz H (ed.),
vol. 1. Academic Press: New York, 1968; 99182.
28. Sneddon IN, Lowengrub M. Crack Problems in the Classical Theory of Elasticity. Wiley: New York, 1969.
29. Irvin GR. Analysis of stresses and strains near the end of a crack traversing a plate. Journal of Applied Mechanics
1957; 24:361364.
30. Garagash DI. Propagation of a plane-strain fluid-driven fracture with a fluid lag: early-time solution. International
Journal of Solids and Structures 2006, in press, doi:10.1016/j.ijsolstr.2005.10.009.
31. Detournay E. Fluid and solid singularities at the tip of a fluid-driven fracture. In IUTAM Symposium on Non-Linear
Singularities in Deformation and Flow, Durban D, Pearson JRA (eds). Kluwer Academic Publishers: Dordrecht,
1999; 2742.
32. Garagash DI, Detournay E. Viscosity-dominated regime of a fluid-driven fracture in an elastic medium. In
IUTAM Symposium on Analytical and Computational Fracture Mechanics of Non-Homogeneous Materials,
Solid Mechanics and Its Applications, Karihaloo BL (ed.), vol. 97. Kluwer Academic Publishers: Dordrecht, 2002;
2529.
33. Adachi JI. Fluid-driven fracture in permeable rock. Ph.D. Thesis, University of Minnesota, Minneapolis, 2001.
34. Rubin AM. Tensile fracture of rock at high confining pressure: implications for dike propagation. Journal of
Geophysical Research 1993; 98:15,91915,935.
35. Khazan YM, Fialko YA. Fracture criteria at the tip of fluid-driven cracks in the earth. Journal of Geophysical
Research 1995; 22:25412544.
Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

PLANE-STRAIN FRACTURE DRIVEN BY A POWER-LAW FLUID

1475

36. Dyskin AV, Germanovich LN. A model of crack growth in microcracked rock. International Journal of Rock
Mechanics and Mining Science and Geomechanics Abstracts 1993; 30:813820.
37. Papanastasiou P. The effective fracture toughness in hydraulic fracturing. International Journal of Fracture 1999;
96:127147.
38. Lecampion B, Detournay E. An implicit algorithm for the propagation of a hydraulic fracture with a fluid lag.
Computer Methods in Applied Mechanics and Engineering 2006, submitted.
39. Kresse O, Detournay E, Garagash DI. Tip solution of a fracture driven by a power law fluid in permeable rock,
2004, unpublished work.

Copyright # 2006 John Wiley & Sons, Ltd.

Int. J. Numer. Anal. Meth. Geomech. 2006; 30:14391475


DOI: 10.1002/nag

You might also like