Sei sulla pagina 1di 31

FL48CH12-Detournay ARI 3 December 2015 9:24

ANNUAL
REVIEWS Further
Click here to view this article's
online features:
• Download figures as PPT slides
• Navigate linked references
• Download citations
• Explore related articles
Mechanics of Hydraulic
• Search keywords
Fractures
Emmanuel Detournay
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Department of Civil, Environmental, and Geo- Engineering, University of Minnesota,


Minneapolis, Minnesota 55455; email: detou001@umn.edu
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Annu. Rev. Fluid Mech. 2016. 48:311–39 Keywords


The Annual Review of Fluid Mechanics is online at multiscale, tip asymptotics, similarity solutions, penny-shaped fracture
fluid.annualreviews.org

This article’s doi: Abstract


10.1146/annurev-fluid-010814-014736
Hydraulic fractures represent a particular class of tensile fractures that prop-
Copyright  c 2016 by Annual Reviews. agate in solid media under pre-existing compressive stresses as a result of
All rights reserved
internal pressurization by an injected viscous fluid. The main application of
engineered hydraulic fractures is the stimulation of oil and gas wells to in-
crease production. Several physical processes affect the propagation of these
fractures, including the flow of viscous fluid, creation of solid surfaces, and
leak-off of fracturing fluid. The interplay and the competition between these
processes lead to multiple length scales and timescales in the system, which
reveal the shifting influence of the far-field stress, viscous dissipation, frac-
ture energy, and leak-off as the fracture propagates.

311
FL48CH12-Detournay ARI 3 December 2015 9:24

1. INTRODUCTION
Hydraulic fractures are tensile cracks that are driven by high-pressure fluid injection. They prop-
agate in solid media subjected to pre-existing compressive stresses. Hydraulic fractures are most
commonly engineered for the stimulation of hydrocarbon-bearing rock strata to increase the pro-
duction of oil and gas wells (Economides & Nolte 2000), but there are other applications, such
as remediation projects in contaminated soils (Murdoch 2002) or preconditioning and cave in-
ducement in mining (van As & Jeffrey 2002, Jeffrey et al. 2013) (see the intersection of a hydraulic
fracture with a tunnel face shown in Figure 1). Furthermore, similar hydraulic fractures manifest
naturally at the geological scale as kilometer-long vertical dikes bringing magma from deep un-
derground chambers to the Earth’s surface (Spence et al. 1987, Lister & Kerr 1991, Rubin 1995,
Roper & Lister 2005, 2007); subhorizontal fractures, known as sills, that divert magma from dikes
(Pollard & Hozlhausen 1979, Spence & Sharp 1985, Emerman et al. 1986, Bunger & Cruden
2011); and cracks propagating at glacier beds (Tsai & Rice 2010).
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Since the pioneering work of Khristianovic & Zheltov (1955) and Barenblatt (1956), there have
been numerous contributions to the modeling of fluid-driven fractures (see Mendelsohn 1984,
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Adachi et al. 2007, and Bunger et al. 2007 for an extensive list of references with a particular focus
on the petroleum industry). The essential complexity of the problem linked to the existence of
a moving boundary and to the degeneracy of the nonlinear equations near the crack tip was not
fully recognized, however, when the first models were being developed. In these early attempts,
analytical solutions for plane-strain and radial hydraulic fractures were built based on ad hoc
assumptions (Geertsma & de Klerk 1969; Abé et al. 1976, 1979; Nilson & Griffiths 1983; Nilson
1986; Advani et al. 1987), and numerical models inherited propagation algorithms from dry cracks
based on linear elastic fracture mechanics (LEFM) (Clifton & Abou-Sayed 1979, Clifton 1989,

Hydraulic fracture
(red plastic proppant)

Borehole

Figure 1
A hydraulic fracture exposed by mining in the tunnel face at the Northparkes Mine E48 orebody at a depth
of 580 m below the surface. Red plastic proppant was added to the water used to form this fracture to allow
mapping of the fracture geometry after mining. The 96-mm-diameter borehole, which provided access to
place the fracture, is visible in the center of the photo. Figure adapted from Jeffrey et al. (2009a).

312 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

Vandamme & Curran 1989, Advani et al. 1990, Sousa et al. 1993, Shah et al. 1997). Nonetheless,
these efforts led to a progressive recognition of the multiscale nature of the hydraulic fracturing
problem, a consequence of the interplay between competing physical processes affecting fracture
propagation—namely, the flow of viscous fracturing fluid, creation of fracture surfaces in the solid,
formation of a lag between the crack edge and the fluid front, elastic deformation of the solid, and
leak-off of fluid from the fracture.
The multiscale nature of the problem transpires both through the presence of multiple relevant
length scales in the tip region and through the existence of multiple timescales in the fracture
evolution. The length scales of the tip asymptotics and the timescales of fracture growth are
entangled, however, as the global dynamics of the crack are intimately coupled across scale to
the asymptotic behavior of the crack tip. The existence of similarity solutions for simple fracture
geometries, which are particular time asymptotics in the solution trajectory, is directly linked to
the dominance of specific singular tip asymptotics.
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

This review provides an introduction to the mechanics of fluid-driven fractures and focuses
particularly on the asymptotic representation of the tip region and on the propagation of a penny-
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

shaped fracture for a constant injection rate, both discussed in the context of relevant length scales
and timescales. The near-tip asymptotics are key to the construction of accurate and efficient
solutions of hydraulic fractures, whereas the penny-shaped fracture serves as a didactic illustration
of the evolving regimes of propagation that are associated with changes in the relative influence of
the controlling parameters with time. The decision to narrow the scope of this review to these key
issues has the unfortunate consequence of neglecting some important topics, such as interaction
between closely spaced fractures, buoyancy-driven fractures, and computational algorithms spe-
cific to fluid-driven fractures—an area of intense current activity. References provided in Section 6
point to works concerning some of the significant issues not covered here.
The review is organized as follows. Sections 2–4 concern hydraulic fractures in impermeable
rocks and provide an analysis of the competing role of far-field stress, rock toughness, and fluid
viscosity on the evolution of a penny-shaped fracture and on the asymptotic solution in the tip
region. Section 5 extends this analysis to cases in which there is leak-off of the fracturing fluid
into permeable rocks. Finally, Section 6 concludes with a survey of outstanding issues warranting
future research.

2. HYDRAULIC FRACTURE IN IMPERMEABLE ROCK

2.1. Planar Crack


This section considers the propagation of a planar fracture driven by injection of a fluid in a rock
medium (Figure 2). A hydraulic fracture is characterized by two distinct fronts moving with time
t: crack edge Cc (t) and fluid front Cf (t), which is contained inside Cc (t). Contour Cc (t) defines the
crack footprint, Ac (t), whereas Cf (t) defines the fluid-filled fracture domain Af (t) ⊆ Ac (t). Under
certain conditions, however, the lag λ between the crack edge and the fluid front can be negligible,
and the two fronts effectively coalesce to a single front, C(t) = Cc (t) = Cf (t). Furthermore, enough
time is presumed to have elapsed since injection began, so that the dimension of the two fronts
is large enough to permit the fluid injection to be viewed as occurring at a point. The injection
point conveniently serves as the origin for position vector x.
Four sets of assumptions are introduced to simplify the problem and make it more tractable;
they pertain to the rock, the fluid, the in situ stress, and the lag region, respectively (Detournay &
Peirce 2014). First, the rock is homogeneous, linearly elastic, brittle, impermeable, and of infinite
extent. Thus, the parameters needed to characterize the rock are Young’s modulus E, Poisson’s

www.annualreviews.org • Mechanics of Hydraulic Fractures 313


FL48CH12-Detournay ARI 3 December 2015 9:24

VC

Vf

Q y

x f c

X f*
S λ

Xc
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

nC
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Figure 2
Planar hydraulic fracture with two distinct moving fronts: crack edge Cc (t) and fluid front Cf (t). Figure
adapted from Detournay & Peirce (2014).

ratio ν, and fracture toughness K Ic , which is a measure of the energy required to create new surfaces
in the rock by the propagation of a tensile crack. Second, the fracturing fluid is incompressible and
Newtonian with viscosity μ. Third, the orientation of the minimum in situ compressive principal
stress, σ 0 (x), is uniform. The assumptions on the far-field stress and on the rock medium ensure
that the fracture propagates in mode I (pure tension). Finally, the lag region between the crack
edge and fluid front is filled with vapors from the fracturing fluid, at a pressure that is negligible
compared to the magnitude of the far-field stress. The hydraulic fracturing problem thus can be
formulated as follows. Given the fluid and solid properties, far-field stress, and injection history,
one can determine how the two fronts Cc (t) and Cf (t), crack opening field w(x, t), and fluid pressure
p f (x, t) evolve from the initial conditions described by Cc (0), Cf (0), w(x, 0), and p f (x, 0).

2.2. Governing Equations


Two fundamental equations govern the fracture opening, w (x, t), and the fluid pressure, p f (x, t)
(or the net pressure, p = p f − σ0 ): the nonlocal elasticity equation and the nonlinear Reynolds
lubrication equation. Together with the boundary conditions on Cc (t) and on Cf (t), and the initial
conditions, they provide a complete formulation of the evolution problem. Three combinations
of material parameters are introduced to simplify these equations:
 1/2
E 2
E = K = 4 K Ic , μ = 12μ, (1)
1 − ν2 π

which are simply referred to as the elastic modulus (E , also known as the plane-strain modulus),
toughness (K ), and viscosity (μ ).

2.2.1. Elasticity. For an elastic domain of infinite extent, the elasticity equation can be expressed
as a hypersingular integral equation (Crouch & Starfield 1983, Hills et al. 1996), given by

E w(x , t)dAc (x )
p(x, t) = p f (x, t) − σ0 (x) = − x ∈ Ac (t). (2)
8π Ac (t) |x − x |3

314 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

This integral equation is obtained by equating the net pressure in the fracture to the normal stress
induced by a continuous distribution of normal dislocation dipoles [nominally characterized by
the opening δ(x), where δ(x) is the Dirac function], with density w(x, t).

2.2.2. Lubrication. The lubrication model describes the flow of viscous fluid in the fracture,
which is slow enough that any inertial effects can be neglected. The lubrication approximation
leads to the nonlinear Reynolds equation (Batchelor 1967),
∂w 1  
=  ∇ · w3 ∇ p f + Q(t)δ(x), x ∈ Af (t), (3)
∂t μ
which is obtained by combining Poiseuille’s law for the fluid flux, q,
w3
q=− ∇ pf , (4)
μ
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

with the continuity equation,


Access provided by 165.22.104.127 on 04/16/20. For personal use only.

∂w
+ ∇ · q = Q(t)δ(x). (5)
∂t
Gravity is neglected in Equation 3 because it has an insignificant role in the class of problems treated
here. The boundary condition at the injection point is incorporated directly via the singular term
Q(t)δ(x).

2.3. Boundary Conditions on the Moving Fronts


The boundary conditions on the moving fronts are formulated first for the general case when
the fluid front lags behind the crack edge, and then for the degenerated case when the two fronts
have coalesced.

2.3.1. Finite fluid lag. There are two boundary conditions on each moving front:

w(xc , t) = 0, K I (xc , t) = K Ic , xc ∈ Cc (t), (6)

p f (xf , t) = 0, Vf (xf ) = q(xf )/w(xf ), xf ∈ Cf (t). (7)

2.3.1.1. Crack front Cc (t). The second condition in Equation 6 is an expression of the critical
equilibrium at each point of the crack front, Cc (t). It reflects the condition that the stress intensity
factor, KI , which is the strength of the stress singularity at the edge of a tensile crack, is everywhere
equal to the toughness (Irwin 1957, Rice 1968). This condition means that in an infinitesimal
advance of the front, the energy flowing to the crack tip is exactly equal to the energy dissipated
in the creation of new surfaces in the solid. Let nc (xc ) denote the outward unit normal to the front
and Vc (xc ) the velocity of the crack front, with Vc = Vc · nc , as Vc is normal to the crack front.
Continuous propagation of the front, implied by writing K I = K Ic everywhere along the crack
front, means therefore that Vc (xc ) > 0. Hence, situations are purposely avoided in which part
of the front either stops moving (Vc = 0) or recedes (Vc < 0). These situations are unlikely to
occur for the class of problems considered in view of the continuing injection, Q(t) > 0, and the
assumed smooth variation of the far-field stress, σ0 (x). Nonetheless, in more general cases that
include shut-in [Q(t) = 0], flowback [Q(t) < 0], leak-off of fluid in the rock, and stress jumps
across layers, the mobile equilibrium condition must be relaxed to allow for 0 ≤ K I ≤ K Ic .

www.annualreviews.org • Mechanics of Hydraulic Fractures 315


FL48CH12-Detournay ARI 3 December 2015 9:24

Both conditions in Equation 6 can be represented by imposing, in the vicinity of Cc (t), the
asymptotic form of the crack opening according to LEFM:
s →0 K  1/2
w(xc − s nc ) ∼ s , (8)
E
where s is the distance of a point x ∈ Ac inside the crack from the crack front. The asymptotic
variation of the crack opening with s1/2 is consistent with the assumed regular net pressure loading
in the lag region (Williams 1952, Kanninen & Popelar 1985).

2.3.1.2. Fluid front Cf (t). The second condition in Equation 7 is a Stefan condition on the fluid
front Cf (t); it expresses that the local filling rate of the crack at the moving front, V f w(xf ), is equal
to the flux, q (xf ), noting also that q(xf ) is orthogonal to Cf (t), as the front corresponds to the
constant pressure boundary condition in Equation 7.
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

2.3.2. No fluid lag. Under certain conditions, which are typical of most hydraulic fracturing
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

treatments, the lag between the fluid front and crack edge is negligible, so the two distinct
fronts effectively coalesce into a combined front, Cc (t) = Cf (t) = C(t), with the consequences that
(a) the pressure boundary condition p = −σ0 (equivalent to p f = 0) disappears as the net pressure
becomes singular, and (b) the Stefan condition wVf = q degenerates to q = 0. There are now
three distinct conditions imposed at the moving front C(t):

w(xc , t) = 0, q(xc , t) = 0, K I (xc , t) = K Ic , xc ∈ C(t). (9)

The degeneracy of the boundary conditions on Cf (t) upon coalescence of the two fronts is un-
derstood best through a limiting process (Detournay & Peirce 2014), after noting that the lag
λ decreases and eventually vanishes as the dimensionless stress σ0 K 2 / μ V E 2 becomes large
(Garagash & Detournay 2000) (see also Section 3). Indeed, letting xf∗ = xc − λ(xc )nc (xc ) denote
the point on Cf (t) closest to xc on Cc (t), one obtains, with vanishing lag λ, xf∗ (xc , t) → xc and
w(xf∗ , t) → 0, whereas Vf (xf∗ ) → Vc (xc ) = V(xc ); thus, q(xf∗ , t) → 0 and p(xf∗ , t) → −∞ when
λ → 0.

2.4. Solving for the Fracture Evolution


The solution can be advanced in time, starting with the initial positions of fronts Cc (0) and Cf (0),
and consistent initial fields w(x, 0) and p(x, 0), by solving the elasticity equation (Equation 2) and
the lubrication equation (Equation 3) with the embedded source condition Q(t) for w(x, t), p(x, t),
Cc (t), and Cf (t) using the boundary conditions in Equations 6 and 7 or 9—depending on whether
Cc (t) and Cf (t) are distinct or have coalesced. Regarding the more challenging zero-lag case, the
following three independent conditions in Equation 9 on C(t) provide the necessary information
to evolve the solution (Detournay & Peirce 2014):
 Given the opening field w(x, t) and domain contour C(t), the condition q = 0, which amounts
to a Neumann boundary condition, is sufficient to solve for the net pressure, p(x, t), up to a
constant.
 Given the net fluid pressure p(x, t) (up to a constant), C(t), and the crack volume Vc (t)
t
[equal to the injected volume of fluid Vf (t) = 0 Q(τ )d τ + Vf (0)], the condition w (xc , t) = 0
is sufficient to solve for opening w(x, t), noting that the undetermined constant in the
fluid pressure is determined by the solvability condition, Vc (t) = Vf (t), typical of Neumann
problems.

316 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

 Propagation criterion K I = K Ic provides the supplementary information to determine the


position of crack edge C(t).
Solving the fracture evolution problem still represents a formidable task, even under the ide-
alized conditions that have been assumed here. The challenges are manifold. In addition to the
challenge of solving the coupled nonlocal and nonlinear system of integrodifferential equations
that governs the flow of a viscous fluid in a deformable channel, there is the inherent difficulty
brought on by the moving boundary nature of the problem. In particular, the difficulty of tracking
the moving front C(t) is compounded by the inability to apply standard computational algorithms
for solving moving boundary problems, such as the volume of fluid method (Voller 2009) or the
level set method (Osher & Sethian 1988, Sethian 1999), as the front velocity, V, does not appear
explicitly in any of the three conditions in Equation 9 on C(t). Furthermore, in the tip region,
application of Poiseuille’s law (Equation 4) and the continuity equation (Equation 5) presents a
challenge; this is well posed analytically, but the numerical evaluation is problematic because this
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

operation entails the product of a vanishing fracture opening squared and a pressure gradient that
tends toward infinity at the front. An alternative, discussed below, is to make explicit use of the
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

tip asymptotic.
There is also the issue of selecting appropriate initial conditions because of the challenging
scale mismatch between the initial phase of fracture initiation at the borehole wall and stages of
propagation when injection can be viewed as occurring at a point. It is neither currently feasible
nor even desirable to conduct computations that span such a huge range of length scales. The
pseudo-initial solution has to correspond to a later time when the details of the early stage of
growth have become inconsequential. Yet this initial solution has to satisfy the governing system
of equations.

3. TIP ASYMPTOTICS IN IMPERMEABLE ROCK

3.1. Governing Equations near an Advancing Tip


In the vicinity of the crack front, the elasticity equation (Equation 2) and lubrication equation
(Equation 3) degenerate into the equations governing the solution for a semi-infinite hydraulic
crack propagating at constant velocity V. Thus, the asymptotic behavior presents a traveling wave
solution. These results imply that, near the crack front, the dependence of the opening and pressure
on the position and time reduces to a dependence on the distance from the crack front, s, with the
current local-front velocity acting as a time-dependent parameter. Assuming that the unknown lag
λ̂ between the crack and fluid front is small compared to the local radius of the crack edge—or is
nonexistent—and that the far-field stress, σ 0 , is uniform at this scale leads to a degenerate system
of equations that govern the asymptotic p̂(s ) and ŵ(s ), where fˆ notation is used to designate the
asymptotics of a quantity f.
The elasticity equation in the tip region is deduced from the planar integral elasticity equation
(Equation 2) by considering distances from the front that are small compared to its local radius of
curvature and by noting that any variation of the crack opening in the close vicinity of the crack
front takes place in the direction normal to the front, as w = 0 on the front (Peirce & Detournay
2008):

E  ∞ dŵ dz
p̂ = , s > 0. (10)
4π 0 dz s − z
Thus, the degenerated elasticity equation in the tip region is asymptotically identical to the singular
integral equation for a semi-infinite crack under plane-strain conditions (Rice 1968). This equation

www.annualreviews.org • Mechanics of Hydraulic Fractures 317


FL48CH12-Detournay ARI 3 December 2015 9:24

accepts a particular class of exact solutions of the form (Williams 1952, Kanninen & Popelar 1985)
ŵα = Cs α , p̂ α = 14 CE cot(π α)s α−1 , 0 < α < 1, (11)
to which particular asymptotics of the tip region are shown to belong.
Because p̂(s ) and ŵ(s ) vary only with the distance from the fracture front, Poiseuille’s law
(Equation 4) simplifies to
ŵ3 d p̂
q̂ = , s > λ̂, (12)
μ ds
where flux q̂ is taken as positive in the direction of negative s. Furthermore, the continuity equation
(Equation 5) reduces to (Desroches et al. 1994)
q̂ = V ŵ, s ≥ λ̂, (13)
which implies that the average fluid velocity in the tip region is equal to the current local velocity
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

of the crack front. This asymptotic form of the continuity equation is obtained in a sequence of
steps that involves (a) expressing the storage term ∂ ŵ/∂t in Equation 5 with regard to the moving
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

coordinate s; (b) recognizing that the convective term V dŵ / ds (where V is the instantaneous
local propagation velocity of the fracture front) dominates the time derivative at a fixed s, when
s → 0 (as can be surmised from the power-law solution w ∼ s α , with α < 1 in accordance with
Equation 11); and (c) integrating the reduced differential equation using the Stefan condition,
q̂ (λ̂) = V ŵ(λ̂), at the moving fluid front, which degenerates to a zero flux condition, q̂ (0) = 0, if
λ̂ = 0. The combination of Equations 12 and 13 finally leads to an asymptotic form of the Reynolds
equation:
ŵ2 d p̂
= V , s > λ̂. (14)
μ ds
Time therefore does not appear explicitly in the system of equations governing the asymptotic
fields of the tip, which depend only parametrically on time via the evolving tip velocity, V.
The elasticity equation (Equation 10) and lubrication equation (Equation 14), together with
the propagation criterion,
ŵ K
lim 1/2 =  , (15)
s →0 s E
and the pressure condition in the lag zone, p̂(s ) = −σ0 (with 0 < s < λ̂ if a lag exists), form a
closed system that can be solved for p̂(s ), ŵ(s ), on the semi-infinite interval [0, ∞), and for lag λ̂.
Although this seems to require a numerical solution, the large s asymptotics can be deduced simply
by combining power-law elastic solutions (Equation 11) and the Reynolds equation (Equation 14)
to yield
  1/3  4 1/3
s →∞ μV s →∞ −2/3 −2/3 E
ŵ ∼ 21/3 35/6 s 2/3
, p̂ ∼ −2 3 s −1/3 . (16)
E μ V
Therefore, the far-field behavior of the system, as represented by Equation 16, depends neither
on toughness K nor on far-field stress σ 0 . Actually, Equation 16 represents the exact solution for
a steadily advancing semi-infinite fracture when λ̂ = 0 and K  = 0 (Desroches et al. 1994); it is re-
ferred to as the viscosity asymptote, or m-asymptote. Originally derived by Spence & Sharp (1985)
as the tip asymptote for a finite hydraulic fracture propagating in a zero-toughness solid (see also
Lister 1990), Equation 16 was later shown to represent an intermediate asymptote for hydraulic
fractures propagating in the viscosity-dominated regime (Garagash & Detournay 2000, 2005).
In summary, the tip asymptotics are given at any time t by the solution of a stationary semi-
infinite crack moving at constant velocity V, equal to the local current tip velocity of a finite crack
at time t. Thus, the tip solution is autonomous.

318 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

103
σ0

102 2
^ ^
pf w s V 3
^
101 λ

100
^ .11
Ωms =4
κ ms
10 –1

10 –2
=0
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

s
m
κ

1
10 –3 2
3
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

2
10 –4 –4
10 10 –3 10 –2 10 –1 100 101 10 2 103
ξms

Figure 3
Crack opening
˜ ms (ξms ) in log-log scale for several values of the dimensionless toughness κ ms (0, 0.05, 0.145,
0.49, 0.99, 2.084, 3.327, and 4.11). The dashed line corresponds to the zero-toughness singular solution.
Figure adapted from Garagash & Detournay (2000).

3.2. Asymptotics with Lag


The introduction of the following scales for length, net pressure, and opening, as well as the
dimensionless toughness, κ ms ,
 
μ VE2 μ VE K σ0 1/2
ms = , p ms = σ 0 , wms = , κms = , (17)
σ03 σ02 E  μ V
into the system of equations governing the near-tip solution indicates that the traveling wave so-
lution is of the form
ˆ ms (ξ̂ms ; κms ) [where

ˆ ms (ξ̂ms ; κms ) and  ˆ ms = ŵ(s )/wms and ˆ ms = p̂(s )/ p ms ]


and that there is a universal relation, ms (κms ), between the lag λ̂ / ms and κ ms (Garagash &
Detournay 2000). Figure 3 presents a log-log plot showing the numerically constructed crack
opening
ˆ ms (ξ̂ms ) for several values of κ ms . It can be observed that the solution behaves according
to the linear elastic fracture asymptotes near the tip (
ˆ ms ξms∼→0 ξ̂ms
1/2
if κms > 0 and
ˆ ms ξms∼→0 ξ̂ms
3/2
if
ξms →∞
κms = 0) and according to the viscosity asymptote far from the tip (
ˆ ms ∼ ξ̂ms
2/3
). In fact, ms rep-
resents an upper bound of the region where the k-asymptote (Equation 15) applies (see Figure 3),
whereas the length scale, mk (defined in Equation 18), provides a measure of the distance from
the tip beyond which the m-asymptote (Equation 16) prevails. Note that toughness κms can be
expressed in terms of these two length scales: κms = ( mk / ms )1/6 .
The lag decreases monotonically with κms [noting that ms (0) 0.36, ms (0.49) 0.2,
ms (0.99) 0.1, and ms (2.1) 0.01] and is given asymptotically by ms = ∗ κms 6
exp(−κms 2
)
−3
with ∗ 4.36 10 for κms  4. These results can be used to assess the influence of the far-
field stress, σ 0 , on the lag by now interpreting the dimensionless toughness as κms = (σ0 /σ∗ )1/2
with σ∗ = μ VE2 /K 2 . For example, taking σ0 = 25 MPa (indicative of a depth of approximately
1,000 m), E  = 3 × 104 MPa, K  = 3 MPa · m1/2 , μ = 10−6 MPa · s (corresponding to a fracturing
fluid viscosity of approximately 100 cp), and the tip velocity V = 1 m/s yields σ∗ = 100 MPa,

www.annualreviews.org • Mechanics of Hydraulic Fractures 319


FL48CH12-Detournay ARI 3 December 2015 9:24

κms = 0.5, and ms 5.8 cm. A lag of λ 1.2 cm is predicted under these conditions; the lag
would reach approximately 2 cm if K  = 0, as would be the case if a pre-existing discontinuity
were reopened by fluid injection. Under a constant injection rate, the tip velocity decreases as the
fracture grows, as a result of the increasing length of the crack front and storage. With time, the
lag is expected to decrease as κms ∼ V −1/2 .
These estimates suggest that under typical conditions, hydraulic fractures induced at depths
of thousands of meters have lags no larger than a few centimeters. Because such a lag is very
small compared to the fracture dimension (which can reach hundreds of meters), it is generally
appropriate to model deep hydraulic fractures as having no lag.

3.3. Asymptotics Without Lag


For large toughness, κms (effectively when κms  4), the tip asymptotics can be written as
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

ˆ ms (ξms ; κms ) = κms


4 ˆ −6 −2 ˆ
ˆ ms (ξms ; κms ) = κms −6


mk (κms ξms ) and  mk (κms ξms ) (Garagash & Detournay
ˆ mk (ξmk ) and
−6
ˆ mk (ξmk ), with ξmk = κms
2000). In other words, the tip solution can be expressed as
ξms
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

and with the new scales for the length, pressure, and opening given by

K 6 μ VE2 K 4
mk = p mk = , wmk = . (18)
E μ2 V 2
4 K 2 μ VE3
ˆ mk (ξmk ), has near-tip and far-tip series expansions for

The rescaled theoretical solution,


ˆ mk (ξmk )
given by
1/2
ˆ mk = ξmk 3/2 3/2
ξmk → 0:
+ 4π ξmk + 64 ξmk ln ξmk + O(ξmk ), (19)

2/3
ˆ mk = β0 ξmk
ξmk → ∞:
+ β1 ξmk
h
+ o (ξmk
h
), (20)

where β0 = βm0 = 21/3 35/6 , β1 0.0371887, and h 0.138673 (Garagash & Detournay 2005).
This large toughness solution,
ˆ mk (ξmk ) and 
ˆ mk (ξmk ), is characterized by a virtual absence of
lag, independent with respect to σ 0 (except as a reference for the definition of the net pressure),
and a logarithmic pressure singularity at the tip,  ˆ mk (ξmk ) ξmk∼→0 ln ξmk . Indeed, such a pressure
singularity for σ0 / σ∗ → ∞ is expected, given that, at the fluid front,  ˆ mk (
ˆ mk ) = −σ0 /σ , with
σ0 /σ∗ → ∞
the lag asymptotically given by  ˆ mk ∼ = ∗ exp (−σ0 /σ∗ ). More precisely, the logarithmic-
pressure singularity results from integrating the lubrication equation,
ˆ mk /dξmk = 1, using
ˆ 2mk d
1/2
the dominant term of the near-field crack opening, ξmk .
Figure 4 plots the theoretical solution,
ˆ mk (ξmk ), together with experimental data obtained in a
series of laboratory hydraulic fracturing experiments (Bunger & Detournay 2008). In these experi-
ments, glycerin or glucose-based solutions were injected into PMMA [poly(methyl methacrylate)]
or glass specimens subjected to confining stresses to propagate penny-shaped cracks. The full-
field fracture opening was assessed by analyzing the loss of intensity as light passes through the
dye-laden fluid that fills the crack (Bunger 2006). The data points represent a composite of 10 ex-
periments that were designed through appropriate selection of the solid (elasticity and toughness),
fluid (viscosity), and injection rate so as to change, via mk , the behavior in the near-tip region un-
der large-enough confining stress, σ 0 , to ensure that the lag is negligible. Comparison of
ˆ mk (ξmk )
with the dominant term of the near- and far-tip expansion (Equations 19 and 20) asymptotes
indicates that the LEFM k-asymptote applies for ξmk  10−5 and the viscosity m-asymptote for
ξmk  10−1 .

320 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

101
0.03
ℓmk /R ≈ 105 k m

w (mm)
0.02

10 0
^
0.01 m 2
3
0
0 0.1 0.2 0.3
s/R

10 –1

^
Ωmk

10 –2

m
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

0.06 ℓmk /R ≈ 1
w (mm) 0.04
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

10 –3 0.02 k
^

Average
0
1 0 0.1 0.2 0.3
s/R
k
2
10 –4 –8
10 10 –6 10 –4 10 –2 10 0
ξmk

Figure 4
Theoretical (black line) mk-solution crack-opening
ˆ mk (ξmk ) with the (blue) m-asymptote and (red )
k-asymptote (Garagash & Detournay 2000), together with data from a series of laboratory hydraulic
fracturing experiments involving the propagation of penny-shaped cracks (Bunger & Detournay 2008). The
solid line (dark yellow) represents the average of all data. One test is shown in gray dots to indicate the portion
of the tip solution exhibited in a typical experiment. The two insets show the crack opening in the tip region:
(top left) a toughness-dominated case for which mk / R ≈ 105 , where R is the fracture radius, and (bottom
right) a viscosity-dominated case for which mk / R ≈ 1. Figure adapted from Bunger & Detournay (2008).

3.4. Applications of Tip Asymptotics for Constructing Solutions


The length scale mk of the autonomous tip sets the mode of dissipation near the front of a finite
hydraulic fracture with no lag. A small-enough mk (compared to a scale, , characteristic of the
finite fracture, such as the local radius of curvature of the crack front or the distance between the
front and point of injection) enables the emergence of the viscosity asymptote in the tip region.
Under these conditions, the solution near the front exhibits a boundary layer structure, with the
k-asymptote as the inner solution and the m-asymptote acting at a distance on the order of mk
from the front. Therefore, the local tip solution viewed at the scale of the finite crack does not
depend on the toughness, as the LEFM asymptotic region is shielded by the intermediate viscous
dissipation asymptote. In other words, the solution at scale is virtually the same as the solution
for zero toughness (in which case the viscosity asymptote is strictly the tip asymptote and not an
intermediate one). In contrast, if mk is large relative to , the asymptote visible at the scale of the
finite crack is the k-asymptote, which reflects energy dissipation in the creation of new surfaces in
the rock. Intermediate values of mk / indicate dissipation in both the fluid and solid and do not
translate into a power-law asymptote at length scale .
In summary, two limiting regimes exist that are characterized by different tip asymptotes at
the fracture scale: the k-asymptote w̃ ∼ s 1/2 in the toughness-dominated regime ( mk / 1)
and the m-asymptote w̃ ∼ s 2/3 in the viscosity-dominated regime ( mk /  1). The behavior of

www.annualreviews.org • Mechanics of Hydraulic Fractures 321


FL48CH12-Detournay ARI 3 December 2015 9:24

the tip region can be captured completely by a stationary solution that is translated and rescaled
according to the tip velocity, which enters the autonomous solution simply as a parameter.
The realization that the global solution depends critically on the details of the near-tip solution
has led to the construction of accurate benchmark solutions for simple hydraulic fracture geome-
tries (with zero lag), plane-strain fracture (Carbonell et al. 1999; Adachi 2001; Adachi & Detournay
2002; Garagash & Detournay 2005; Garagash 2006a,c; Peirce 2010), and penny-shaped fracture
(Savitski & Detournay 2002, Madyarova 2004). It also has led to the development of computational
algorithms capable of achieving an accurate solution on a relatively coarse mesh, as demonstrated
in a comparison study of the performance of various numerical simulators of hydraulic fractures
(Lecampion et al. 2013). These new algorithms adapt the asymptotic conditions to the spatial
resolution of the mesh and select the appropriate asymptotic behavior as part of the solution (see,
e.g., Peirce & Detournay 2008 and Peirce 2015 for a description of a fixed-mesh algorithm for
planar hydraulic fractures). Simulation of a fracture propagating in a layer sandwiched between
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

two layers subjected to larger compressive stresses illustrates the ability of such algorithms to
select the relevant asymptotic behavior at any point of the crack front as a function of the local tip
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

velocity (Peirce 2015).

4. PENNY-SHAPED HYDRAULIC FRACTURE IN IMPERMEABLE ROCK

4.1. A Two-Timescale Problem


A fracture evolves as a penny-shaped crack provided that the material properties and initial stress
are uniform over a length scale large enough compared to the ultimate fracture dimension. Under
these conditions, solving the fracture evolution problem is reduced to determining the net pressure,
p(r, t) = p f (r, t) − σ0 , and fracture opening, w(r, t), as a function of radial coordinate r and time
t, as well as the fracture radius R(t) and fluid-front radius Rf (t) [or fluid lag λ(t) = R(t) − Rf (t)]
(see Figure 5). Considerations are restricted here to a constant injection rate, Q0 , as it enables
clarification of the dependence of the fracture evolution on material constants E , K , and μ , as
well as on σ 0 and Q0 .
A scaling analysis reveals that the propagation of a penny-shaped fracture in an impermeable
medium is characterized by two timescales, tom and tmk , defined as (Savitski & Detournay 2002,

σ0 σ0
=0
Q0 Q0

w φ = φ1
pf r O =∞ pf r w

Rf λ R
R =0
φ << 1

Figure 5
Parametric space OMK with the three vertices representing the small-time (O), intermediate-time (M), and large-time (K) similarity
solutions for a penny-shaped hydraulic fracture propagating in an impermeable elastic medium. The evolution of the fracture from the
O- to the K-vertex depends on the timescale ratio ϕ = to m /tmk . The two sketches illustrate a hydraulic fracture near the O-vertex (large
lag) and near the MK-edge (zero lag). Figure adapted from Bunger & Detournay (2007).

322 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

Bunger & Detournay 2007)


 1/2
E 2 μ E 13 μ5 Q 3o
to m = , tmk = . (21)
σo3 K 18
With time measured relative to both timescales, there exists not only (a) a small-time asymptotic
solution, denoted as the O-solution, for t / to m  1 and t / tmk  1, and (b) a large-time asymptote,
the K-solution, for t / tmk 1, but also (c) an intermediate-time asymptote when t / to m 1
and t / tmk  1. This particular solution, denoted as the M-solution, emerges provided there is a
separation of the two timescales (i.e., if to m / tmk <
<< 1).
The time asymptotics are similarity solutions with a power-law dependence on time of the
fluid and crack-front radii, net pressure, and opening (Savitski & Detournay 2002, Bunger &
Detournay 2007). However, they also can be interpreted as particular solutions that correspond
to degenerated values (0 or ∞) of some parameters. This dual interpretation of the similarity
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

solutions is facilitated by first recasting the two times t / to m and t / tmk as


 1/3  1/9
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

t t
Z= , K= (22)
to m tmk
and noting that Z(t) ∼ σ0 can be interpreted as a dimensionless far-field stress, and K(t) ∼ K 
as a dimensionless toughness. The three similarity solutions can then be interpreted as follows.
First, for the O-solution (K = 0, Z = 0), only a small region around the injection point is filled
with fluid (i.e., Rf / R  1); thus, loading at the scale of the fracture can be subsumed to a force
dipole acting at r = 0. The fluid fraction grows with time, however, as R ∼ t 10/27 and Rf ∼ t 12/27
(Bunger & Detournay 2007). Second, for the M-solution (K = 0, Z = ∞), this similarity solution
corresponds to the case of zero toughness and infinite stress (Savitski & Detournay 2002). It is
characterized by the absence of a lag, as the mobile equilibrium of the crack cannot be maintained
with infinite stress acting across a finite lag. Finally, the K-solution (K = ∞) corresponds to the
case of an inviscid fluid and is characterized by a uniform pressure (Savitski & Detournay 2002).
The fracture is fully filled with the fluid, as a lag cannot be sustained with an inviscid fluid. This
physical constraint entails that the K-solution is a large-time asymptotic with respect to tmk only,
irrespective of tom .
At this point, it is useful to introduce the OMK-triangle as a conceptual representation of the
solution space in terms of Z and K (see Figure 5). In this space, the solution evolves from the
O- to the K-vertex along a trajectory characterized by a constant ϕ = to m / tmk , with the fluid front
progressively catching up with the crack front. With the O-, M-, and K-solutions mapped on the
corresponding vertices of this triangle, the OM-edge is associated with K = 0, the OK-edge with
Z = 0, and the MK-edge with Z = ∞. Along each edge, the solution evolves with respect to one
timescale only—namely, with Z along OM and with K along MK and OK. The K-vertex actually
represents a degenerated edge, K = ∞, because the solution does not depend on Z if K = ∞.
There are two particular trajectories linking the O- to the K-vertex: ϕ = 0, consisting of the union
of two edges OM and MK, and ϕ = ∞, corresponding to the edge OK. The OMK representation,
which associates vertices to similarity solutions (which do not have characteristic times) and edges
to segments of a trajectory (which depend on only one characteristic time), enables the visualization
of the changing dependence of the solution on the original physical parameters.
Let us consider, for example, a trajectory ϕ  1, characterized by the separation of the two
timescales to m and tmk . This trajectory brings the solution from the O- to the K-vertex via the
M-vertex. Along the OM-edge, the behavior does not depend on the toughness; that is, as the
hydraulic fracture evolves with increasing Z (which can be interpreted as an increasing far-field
stress), it behaves as if toughness K  = 0. Approaching the M-vertex along OM, the relative lag

www.annualreviews.org • Mechanics of Hydraulic Fractures 323


FL48CH12-Detournay ARI 3 December 2015 9:24

decreases as λ/R ∼ Z −3 ; this can be deduced readily from Equation 17, giving the scale ms for the
lag when K  = 0, after noting that the tip velocity V ∼ R/t because of the power-law dependence
of the fracture radius R on time t at the M-vertex. As the crack evolves on the MK-edge (Z = ∞)
with increasing K, the solution does not depend on the stress σ0 (except as a reference for the
definition of the net pressure), as the lag is zero. Moreover, as K → ∞, the behavior no longer
depends on the fluid viscosity. In summary, the fracture response is independent of σ 0 in the close
proximity of the OK- and the MK-edges, independent of K near the OM-edge, and independent
of μ near the K-vertex. The behavior is said to be viscosity dominated near the M-vertex and
toughness dominated near the K-vertex.
These considerations have practical implications for hydraulic fracturing stimulations of deep
oil and gas reservoirs, as many treatments are characterized by ϕ  1 and by a timescale to m
that is typically several orders of magnitude smaller than the treatment time (Bunger et al. 2007,
Detournay et al. 2007). In such cases, the O-solution is not relevant, as other processes taking
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

place at timescales close to tom have been ignored [e.g., fracture initiation at the borehole taking
into account the hydraulic compliance of the injection system (Lhomme et al. 2005)], and the
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

M-solution actually becomes the appropriate small-time asymptotic solution (noting also that its
similarity solution status implies that prior history has no bearing on the evolution of the fracture
past the M-vertex, should this vertex indeed be on the solution trajectory). Nevertheless, there
are situations in which a fluid lag could have a significant finite size, as sometimes observed in
laboratory experiments (Medlin & Masse 1984, van Dam et al. 1999, Bunger et al. 2005b,c). Within
the context of engineered hydraulic fractures, these situations arise in low-stress environments
(e.g., when the fracture is close to a free surface) or in overpressured reservoirs, when the rock
is permeable enough that the pore fluid filling the lag region is approximately equal to the far-
field reservoir pressure, p 0 (Detournay & Garagash 2003)—in which case, σ0 must be replaced by
σ0 − p 0 in the expression for to m in Equation 21.

4.2. Scaling Along the MK-Edge


Introduction of time-dependent scales for the radius, net pressure, and opening of a penny-shaped
hydraulic fracture evolving along the MK-edge reveals the presence of four dimensionless groups
in the scaled equations. Eliminating three groups by setting their value to unity defines the scales,
with the remaining group identified as an evolution variable. Two physically meaningful scalings
are defined, a viscosity and a toughness scaling, which are associated with different power-law
dependence on time of the scales.

4.2.1. General form. Considering only zero-lag solutions along the MK-edge (which are inde-
pendent of σ0 ), a methodology is now outlined to identify the nature of the self-similar M- and
K-solutions through a scaling analysis. One formally starts by introducing three time-dependent
scales: L(t) for the fracture radius, P(t) for the net pressure, and W(t) for the fracture opening.
These scales, which still need to be identified, can then be used to define fracture radius γ , net
pressure , and opening
, all dimensionless quantities:

R(t) = γ (P(t)) L(t), p = P(t) (ρ, P(t)) , w = W (t)


(ρ, P(t)) . (23)

With these definitions, the dimensionless radial coordinate ρ = r / R(t) (0 ≤ ρ ≤ 1) and a


dimensionless history parameter P(t), which depends monotonically on t, are also introduced.
[The existence of only one dimensionless group, P(t), can be confirmed a posteriori.] Under the
scaling in Equation 23 and the assumptions of axisymmetry, no lag, and constant injection rate,

324 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

the general set of Equations 2–5 and 9 governing the propagation of a planar hydraulic fracture
can be simplified as follows.

4.2.1.1. Elasticity. The singular integral equation (Equation 2) can be transformed after integra-
tion (Hills et al. 1996, Gordeliy & Detournay 2011a):

Ge 1 ∂

=− G(ρ, η) d η, (24)


γ 0 ∂η
where kernel G(ρ, η) is given by
⎧  2  2

⎪ 1 η ρ η
1 ⎨ K + E , ρ>η
G(ρ, η) = ρ ρ 2 η 2 − ρ2
 2 ρ 2
, (25)
2π ⎪
⎪ η ρ
⎩ E , ρ≤η
η2 − ρ 2 η2
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

with K(·) and E(·) denoting the complete elliptic integrals of the first and second kind, respectively
(Abramowitz & Stegun 1972). Kernel G(ρ, η) represents the normal stress across the fracture
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

plane that is induced at ρ by a ring dislocation located at distance η from the axis of symmetry.

4.2.1.2. Lubrication. With the conversion from a fixed (r) to a moving (ρ) coordinate in the
evaluation of the storage term also accounted for, the Reynolds equation (Equation 3) transforms
as
   
1 ∂ 3 ∂ 2 Ẇt L̇t ∂

ρ dγ ∂

ρ
= Gm γ
− ρ + Ṗt − . (26)
ρ ∂ρ ∂ρ W L ∂ρ ∂P γ dP ∂ρ

4.2.1.3. Propagation criterion. By expressing the stress intensity factor as a weighted integral of
the net pressure (Rice 1972), one can write the condition of continuous mobile equilibrium in
Equation 8 as

27/2 1/2 1 
γ  ρ dρ = Gk . (27)
π 0 1 − ρ2
The kernel represents the stress intensity factor induced by a ring force dipole located at ρ, with
its square-root singularity emphasizing the dominant influence of the near-tip crack loading.

4.2.1.4. Global fluid volume balance. The zero-flux condition at the tip and the source condition
are implicitly accounted for when equating the volume of injected fluid to the crack volume
 1
2π γ 2
ρ dρ = Gv . (28)
0

Four dimensionless groups appear in the above equations:


WE Q0 t μ L2 K
Ge = , Gv = 2
, Gm = , Gk = , (29)
LP WL P W 2t PL1/2
where subscripts e, v, m, and k stand for elasticity, volume, viscosity, and toughness, respectively.
The scales L(t), P(t), and W(t) thus can be determined by imposing the values of three of these
groups. Groups Ge and Gv are set naturally to 1. Indeed, elementary elasticity considerations
indicate that the average opening scaled by the fracture length is of the same order as the average
net pressure scaled by the elastic modulus, E ; hence, Ge = 1 implies that
and  are of the same
order. Furthermore, Gv = 1 indicates that the scaled fracture volume is on the order of 1, as it is an
approximate statement of the balance between the fracture volume and volume of fluid injected.

www.annualreviews.org • Mechanics of Hydraulic Fractures 325


FL48CH12-Detournay ARI 3 December 2015 9:24

Different scaling emerges depending on whether Gm = 1 or Gk = 1, with the remaining group


identified with the parameter of interest, P.

4.2.2. Viscosity and toughness scaling. Two alternate scalings can thus be defined: a viscosity
scaling (Gm = 1), with P interpreted as a dimensionless time-dependent toughness, K ≡ Gk , and
a toughness scaling (Gk = 1), with P corresponding to a dimensionless time-dependent viscosity,
M ≡ Gm :
 1/18  1/9  3 13 1/5  2/5
t2 t Q0 E tmk
K = K = , M = μ 
= , (30)
μ5 Q 30 E 13 tmk K 18 t 2 t

noting the inverse relation M = K−18/5 . Scales L(t), P(t) and W(t) are explicitly given by
  3 4 1/9   2 1/3  2 3 1/3
E Q0 t μE μ Q0 t
Lm = , P = , W = , (31)
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

m m
μ t E 2

 1/5  1/5  1/5


Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Q 20 E 2 t 2 K 6 Q 0 K 4 t
Lk = , Pk = , Wk = , (32)
K 2 E  Q0 t E 4

Lm Pm Wm
= K2/5 = M1/9 , = K−6/5 = M−1/3 , = K−4/5 = M−2/9 , (33)
Lk Pk Wk
where the subscripts m and k denote quantities defined in the viscosity and toughness scaling,
respectively.
The solution is thus of the form Fm (ρ, K) in the viscosity scaling and Fk (ρ, M) in the toughness
scaling (with F denoting the solution {γ , ,
}). Two particular solutions can be defined: the
zero-toughness solution, Fmo (ρ) ≡ Fm (ρ, 0), and the zero-viscosity solution, Fko (ρ) ≡ Fk (ρ, 0).
Examination of Equations 24–28 confirms that there is no degeneration of the equations when
dimensionless group P (either K or M) goes to zero in either scaling. Furthermore, these equations
indicate that γmo , <mo>, and <
mo> (i.e., the fracture radius, the mean net pressure, and the mean
opening in the viscosity scaling when K = 0) as well as γko , <
ko>, and <ko> (characterizing the
solution when M = 0) are all on the order of 1. Therefore, the scales provide a good measure of
the zero-toughness and zero-viscosity solutions. The two particular solutions K = 0 and M = 0
(K = ∞) are evidently self-similar and thus do not require an initial condition. This analysis
confirms that the solution does evolve from the viscosity-dominated regime (K  1) toward the
toughness-dominated regime (K 1), as K is a monotonically increasing function of time, and
the transition between viscosity- and toughness-dominated regimes can be understood in terms
of only one parameter, dimensionless time t/tmk .

4.3. Small-Time, Large-Time, and Transient Solutions


The viscosity and toughness scalings lead to the identification of two similarity solutions for a
penny-shaped hydraulic fracture with no lag. These solutions, which exhibit a power-law depen-
dence on time, actually represent the small- and large-time asymptotics for a fracture evolving
along the MK-edge. The transient solution is obtained numerically.

4.3.1. Small-time similarity solution. The similarity M-solution can be constructed from the
system of Equations 24–28 with Ge = Gv = Gm = 1 and Gk = 0 using a method inspired by the
scheme described by Spence & Sharp (1985) for solving the plane-strain hydraulic fracture for
K  > 0 (Savitski & Detournay 2002). The m-asymptote (Equation 16), originally derived as a

326 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

traveling wave–type solution, can be translated readily as an asymptotic for


mo and mo :
ρ→1 ρ→1

mo ∼ 2 · 31/6 γmo (1 − ρ)2/3 , mo ∼ −3−4/3 (1 − ρ)−1/3 . (34)

A first-order approximation of the M-solution Fmo (ρ), computed by keeping only the first term
in a Jacobi polynomial expansion, is given by

γmo 0.6955, mo −A1 (1 − ρ)−1/3 + A2 − A3 ln ρ,


(35)

mo (B1 + B2 ρ)(1 − ρ)2/3 + B3 [(1 − ρ 2 )1/2 − ρ arccos ρ],
with A1 0.2387, A2 0.8593, A3 0.09269, B1 1.034, B2 0.6378, and B3 0.1642
(Savitski & Detournay 2002). This first-order estimate of the opening at the injection point differs
by approximately 0.5% from an approximation built with four terms in the Jacobi polynomial
expansion, noting that the m-asymptote applies to approximately 10% of the fracture radius from
the tip.
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

4.3.2. Large-time similarity solution. The similarity K-solution corresponds to the solution for
an ideal inviscid fluid (μ = 0). The governing equations are again Equations 24–28, but now with
Ge = Gv = Gk = 1, Gm = 0. Because the pressure is uniform, the solution for the opening is given
by Sneddon (1946), with the constant scaled pressure and radius deduced from the propagation
criterion in Equation 27 and the global balance equation (Equation 28) (Abé et al. 1976, Savitski
& Detournay 2002):
 2/5  1/5
3 π  π 1/5 3
γko = √ , ko = ,
ko = (1 − ρ 2 )1/2 . (36)
π 2 8 12 8π
Capturing the effect of small viscosity M  1, which introduces a logarithmic singularity to the
pressure at the source and at the tip, requires a regular asymptotic expansion of the solution to
the order of M in the form Fk (ρ, M) = Fko (ρ) + M Fk1 (ρ) + O(M2 ), following the approach
introduced by Garagash (2000, 2006a) for the plane-strain case (see Savitski & Detournay 2002
for details). Consideration of this viscosity correction indicates that the fracture radius is within
1% of the K-solution for K 3.6 (corresponding to t / tmk 105 ).

4.3.3. Transient solution. To calculate the transition from the viscosity-dominated to the
toughness-dominated regime, we choose a time scaling with fixed characteristic radius Lmk , pres-
sure Pmk , and opening W mk . The timescale is naturally selected to be tmk , given by Equation 21,
as it is the time at which the characteristic quantities become equal in the toughness and viscosity
scalings [i.e., Lmk = Lm (tmk ) = Lk (tmk ), Pmk = Pkm (tmk ) = Pk (tmk ), and W mk = W m (tmk ) = W k (tmk )]
(see Equation 33). The system of equations, deduced from Equations 24–28, is solved naturally on
a moving grid with a fixed number of elements. Madyarova (2004) has described such an algorithm
based on a finite-difference discretization of the Reynolds equation; on the displacement discon-
tinuity method for solving the elasticity equation (Crouch & Starfield 1983), in which the crack
is discretized into a number of concentric dislocation rings; and on the multiscale tip asymptotic
in the tip region. The algorithm marches in time starting with the M-solution; the accuracy of
the algorithm can be assessed by checking that the numerical solution actually evolves toward the
K-solution at large time.
Figure 6 plots the evolution fracture opening at the inlet of the fracture radius, together with
the small- and large-time power-law asymptotics. The figure suggests that the viscosity-dominated
regime extends up to approximately t/tmk ∼ 0.1 (K ∼ 0.77) and that the toughness-dominated
regime starts at approximately t/tmk ∼ 105 (K > 3.6).

www.annualreviews.org • Mechanics of Hydraulic Fractures 327


FL48CH12-Detournay ARI 3 December 2015 9:24

104
a b
102 =0 =0
2 101
1

Ω (0,τ)
5
γ (τ)

9 5
10 0 9
4 1
10 0
=0 =0
10 –2

10 –4 10 –1
10 –4 10 0 104 10 8 10 –4 10 0 104 10 8
τ τ

Figure 6
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Evolution of a penny-shaped fracture along the MK-edge: (a) fracture radius γ = R / Lmk and (b) opening
= w/W mk at the injection
point as a function of τ = t / tmk . The numerical results ( gray line) are shown together with the small- and large-time asymptotes (blue
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

and red lines, respectively). Figure adapted from Madyarova (2004).

4.4. Discussion
The evolution of the global solution for a radial hydraulic fracture with no lag is driven entirely
by the simultaneous decrease of the tip velocity and increase of the crack-front length as time
proceeds. The combined effects cause a progressive shift of the main source of energy dissipation
from viscous flow at small time to fracture energy at large time. This transition is effected by a
change of the dominant asymptote at the fracture scale, from the m-asymptote at small time to
the k-asymptote at large time.
The relative thickness of the boundary layer, at which the transition between the 1/2- and
the 2/3-asymptote takes place in the viscosity-dominated regime, can be expressed in terms of
the dimensionless toughness K. Converting the tip velocity V to Ṙ in the expression for mk in
Equation 18 and assuming that R evolves according to the M-solution leads to mk /R ∼ 10 K6 .
This relation confirms that fracture propagation is in the viscosity-dominated regime as long as
K  0.7, as this upper bound guarantees that mk /R  1, a condition that enables the m-asymptote
to emerge at intermediate distances from the tip that are still small enough relative to the crack
radius. Near the K-vertex, mk is several orders of magnitude larger than R ( mk / R ∼ 10 K36/5 ) and
is therefore unphysical. Although the nature of energy dissipation in the tip region is fundamentally
controlled by the local tip velocity, it can be related in the case of simple fracture geometry, such
as the penny shape, to a global parameter (here K), a function of the injection rate, Q0 , and time.
Similar results can be derived for a plane-strain hydraulic fracture (Carbonell et al. 1999,
Garagash & Detournay 2005, Garagash 2006a), with one important difference. The dimensionless
toughness K = K  (E 3 μ Q o )−1/4 is now time independent, a corollary of the different dimension
of Q0 in two-dimensional (2D) and 3D problems. Consequently, the MK-edge, as well as the
OK-edge (Garagash 2006b), is a locus of similarity solutions. In other words, the partitioning of
energy dissipation between a Newtonian fluid and an elastic brittle solid does not evolve with time
in a plane-strain hydraulic fracture for a constant injection rate.
The basic result from scaling analysis that the propagation regime of a penny-shaped fracture
depends on time scaled by characteristic time tmk has practical consequences for the modeling
and interpretation of hydraulic fracturing treatments, as well as for the design of laboratory ex-
periments. Although this simple model lacks critical components for designing a treatment, such
as leak-off, the presence of stress barriers, and non-Newtonian fluid rheology, it nonetheless can

328 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

be used to identify the key parameters affecting fracture growth under conditions in which the
fracture geometry is not too dissimilar from a circular shape. Consider, for example, this plausible
set of parameters: Q 0 = 0.04 m3 /s, E = 8 GPa, ν = 0.2, K Ic = 1 MPa·m1/2 , and μ = 0.1 Pa·s,
with σ 0 assumed to be large enough for the lag to be negligible. Because the timescale for the MK-
transition is tmk 3 h for this set of parameters, the fracture would propagate over a significant
part of the treatment time—on the order of 1 h, in the viscosity-dominated regime. However, a
laboratory experiment under the same conditions, except for an injection rate that is three orders
of magnitude smaller, would be mostly in the toughness regime, as now tmk 0.4 s. The fracturing
fluid would have to be 10–100 times more viscous to enable observation of the viscosity-dominated
regime in laboratory conditions.

4.5. Power-Law Fluids


Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

With few exceptions, fracturing fluids used to stimulate oil and gas wells are not Newtonian
(Cameron & Prud’homme 1989, Constien et al. 2000), but they can be approximately described
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

as power-law fluids. For such a rheology, Poiseuille’s law (Equation 4) becomes


w2n+1
|q |n−1 q = − ∇ p, M  = 2n+1 (2n + 1)n n−n M , (37)
M

where M is a modified consistency index, and n is a fluid behavior index (n = 1 corresponds to
a Newtonian fluid, with M  = μ ). Most of the fracturing fluids used in the petroleum industry
can be described as shear thinning (0 < n < 1) because their apparent viscosity decreases with the
shear strain rate (Cameron & Prud’homme 1989).
All the solutions obtained for hydraulic fractures with a Newtonian fluid can actually be gen-
eralized for power-law fluids. For example, the tip asymptotics at the M-vertex behave according
s →0 s →0 s →0
to w ∼ s 2/n+2 , p ∼ s −n/n+2 (with p ∼ ln s for a perfectly plastic fluid n = 0). The different
dimension of M compared to μ affects the fracture evolution, as can be deduced from a scaling
analysis. For a plane-strain fracture with a constant injection rate, the solution evolves from the
K- to the M-vertex (Garagash 2006c). (The similarity solution at the M-vertex is described in
Adachi & Detournay 2002.) However, for a penny-shaped fracture with a constant injection rate,
the solution evolves from the M- to the K-vertex if n > 0.5, but from the K- to the M-vertex if
0 ≤ n < 0.5; the MK-edge is a locus of similarity solutions for n = 0.5 (as is the case with a
plane-strain hydraulic fracture for n = 1).

5. HYDRAULIC FRACTURE IN PERMEABLE ROCKS

5.1. Carter’s Leak-Off Model


In hydraulic fracturing treatments of conventional oil and gas reservoirs, fluid loss from the fracture
is considered to be a 1D, pressure-independent process. This infiltration model, usually referred
to as Carter’s leak-off (Carter 1957), is expressed as
C
g(x, t) =  , (38)
t − t0 (x)
where g is the infiltration rate from both fracture faces, t0 (x) is the time at which the crack
front (assumed here to be the same as the fluid front) passes through point x, and C  = 2CL ,
with CL denoting the usual Carter’s leak-off coefficient (in units of L T−1/2 ). The usual assump-
tions leading to Carter’s model are that (a) the fracturing fluid deposits a thin layer of relatively

www.annualreviews.org • Mechanics of Hydraulic Fractures 329


FL48CH12-Detournay ARI 3 December 2015 9:24

low permeability known as the filter cake (usually a mixture of dense polymers, chemical ad-
ditives, and particulates) on the inner faces of the fracture, with a deposition rate proportional
to the leak-off rate, and (b) the filtrate (the fraction of fracturing fluid that percolates through
the filter cake) has enough viscosity to fully displace the fluid already present in the rock pores
(Settari 1985). The inverse square-root dependence on time of the leak-off rate actually arises
as the solution of a 1D diffusion problem. With leak-off, the continuity equation in Equation 5
becomes
∂w
+ ∇ · q + g = Q(t)δ(x). (39)
∂t
As shown below, Carter’s leak-off introduces an additional length scale in the tip asymptote and
an additional timescale in the fracture evolution.
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

5.2. Tip Asymptotes


Tip asymptotics for a hydraulic fracture propagating in a permeable rock can be constructed again
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

using the device of a semi-infinite fracture steadily moving at velocity V in an elastic, permeable
solid under the condition of plane strain (Lenoach 1995, Adachi & Detournay 2008, Garagash
et al. 2011). Here, we consider the zero-lag case. With Carter’s leak-off, the continuity equation
in Equation 13 generalizes to
q̂ = Vŵ + 2 C  V 1/2 1/2
s . (40)
First consider the zero-toughness case K  = 0. With Poiseuille’s law (Equation 12) and
Equation 40 combined, another power-law asymptote, denoted as the m̃-asymptote, emerges
from the class of elastic solutions in Equation 11 (Lenoach 1995):
 3/8  1/8
s →0 μ2 C 2 V s →0 E 10
ŵ ∼ 4.26 s 5/8
, p̂ ∼ −0.138 s −3/8 . (41)
E 2 μ2 C 2 V
s →0
This asymptote is distinguished by a square-root singularity of the leak-off rate (g ∼ s −1/2 ).
The complete traveling wave solution for K  = 0 has been constructed numerically (Adachi &
Detournay 2008), in an approach similar to that used to compute the mk-solution presented in
Section 3.3. This solution, denoted as the mm̃-solution, has a universal form, provided that the
distance from the tip is scaled by mm̃ , defined as

26 C 6 E 2
mm̃ = . (42)
μ2 V 5
The mm̃-solution behaves according to the m̃-asymptote (Equation 41) near the tip, s / mm̃  1,
and according to the m-asymptote (Equation 16) far from the tip, s / mm̃ 1.
Generalized tip asymptotics for a hydraulic fracture (K  > 0, μ > 0,C  > 0) have a complex
multiscale structure arising from the two length scales mk and mm̃ , as defined in Equations 18
and 42 (Garagash et al. 2011). The structure of this generalized asymptote is a function of number
χ:
 
mm̃ 1/6 2C  E 
χ= =  1/2 . (43)
mk KV
The traveling wave solution has been constructed numerically as a function of χ ; it behaves
according to the k-asymptote near the tip and according to the m-asymptote far from the tip. If
χ  50, the m̃-asymptote emerges as an intermediate asymptote (Garagash et al. 2011).

330 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

Table 1 Scales L, P, and W and evolution parameters P 1 and P 2 in the M- (storage-viscosity), K- (storage-toughness), M̃-
(leak-off-viscosity), and K̃ -scaling (leak-off-toughness)
Scaling L P W P1 P2
M  1/9  1/3  1/9  1/18  1/18
E  Q 30 t 4 E 2 μ μ2 Q 30 t K 18 t 2
Km = C 18 E 4 t 7
μ t
E 2 μ5 E 13 Q 30 Cm =
μ4 Q 60
   1/5  1/5  1/5  1/10
K E  Q 0 t 2/5 K 6 K 4 Q 0 t μ5 Q 30 E 13 C 10 E 8 t 3
K E  Q0 t E 4 Mk = K 18 t 2
Ck =
K 8 Q 20

M̃  1/4  1/16  1/16  1/16  1/16


Q 20 t E 12 μ4 C 6 μ4 Q 60 t K 16 t μ4 Q 60
C 2 E 4 C 2
Km̃ = Sm̃ = E 4 C 18 t 7
Q 20 t 3 E 12 μ4 C 2 Q 20

K̃  1/4  1/8  1/8  1/4  1/8


Q 20 t K 8 C 2 K 8 Q 20 t μ4 E 12 C 2 Q 20 K 8 Q 20
C 2 E 8 C 2
Mk̃ = K 16 t
Sk̃ = E 8 C 10 t 3
Q 20 t
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

5.3. Penny-Shaped Fracture


Access provided by 165.22.104.127 on 04/16/20. For personal use only.

The problem of a penny-shaped fracture driven by injection of a fluid at a constant rate Q0 is now
re-examined taking into account fluid loss into the host rock and assuming no lag. In addition to
the two competing dissipative processes associated with the fluid viscosity and solid toughness,
the evolution of this fracture is governed also by competition between fluid storage in the fracture
and leak-off in the surrounding rock to balance the injected fluid. Thus, four primary asymptotic
regimes can be identified, corresponding to each combination of one dominant dissipative process
and one dominant fluid-storage mechanism: the storage viscosity (M), storage toughness (K), leak-
off viscosity ( M̃ ), and leak-off toughness ( K̃ ). The previous M-solution now corresponds to the
zero-toughness, zero-leak-off limiting solution, K  = C  = 0, and the K-solution to μ = C  = 0.
The fracture evolution is mediated by two independent timescales, which can be defined as
(Madyarova 2004)
 5 13 3 1/2  4 6 1/7
μ E Qo μ Qo
tmk = , tm m̃ = . (44)
K 18 E 4 C 18
Timescale tmk , previously defined in Section 4.1, is a measure of the time required to evolve from the
viscosity- to the toughness-dominated regime in the absence of leak-off, whereas tmm̃ characterizes
the transition between the storage- and the leak-off-dominated regime for zero toughness. The
expression for the two independent timescales can be deduced by following the steps outlined
in Section 4.2.1, starting with the general scaling in Equation 23, by considering two evolution
parameters, P1 (t) and P2 (t), instead of one. Thus, the dimensionless radius, net pressure, and
opening have the functional forms γ (P1 , P2 ), (ρ; P1 , P2 ), and
(ρ; P1 , P2 ), respectively. Four
different scalings can be defined in connection with the four primary limiting cases introduced
above, each associated with a specific power-law dependence on time t of scales L, P, and W and
of parameters P1 and P2 (see Table 1). It can be observed that the evolution parameters can take
the meaning of a toughness (Km , Km̃ ), viscosity (Mk , Mk̃ ), storage (Sm̃ , Sk̃ ), or leak-off coefficient
(Cm ,Ck ).
The solution regimes can be conceptualized in the rectangular phase diagram MK K̃ M̃ shown
in Figure 7a. Such a space provides a means to organize the solutions and to understand the
influence of the various parameters on the system response. The four primary regimes (M, K, M̃ ,
and K̃ ) of hydraulic fracture propagation are identified with the vertices of the MK K̃ M̃-space;
they are characterized by only one dissipative process (P1 = 0) and only one component of the
fluid global balance (P2 = 0) (see Table 1). These vertex solutions are similarity solutions, with a
power-law dependence on time of the length, net pressure, and width. In particular, the fracture

www.annualreviews.org • Mechanics of Hydraulic Fractures 331


FL48CH12-Detournay ARI 3 December 2015 9:24

~ ~
~ m k ~
M K 102
~
m
a ~
k
b ϕ = 10 –4
101 e
dg
-e
MK
ϕ1 10 0 4
10
ϕ=
tmm~ γ
ϕ2 10 –1

9
10 –2
m k 4
ϕ3
10 –3
M K
m k 10 –7 10 –5 10 –3 10 –1 101 103 105 107
tmk τ
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Figure 7
(a) Parameter space MK K̃ M̃ for a zero-lag penny-shaped hydraulic fracture propagating in a permeable rock. The system evolves with
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

time along a φ -trajectory, starting from the M-vertex (viscosity-storage-dominated regime) and ending at the K̃ -vertex (toughness-leak-
off-dominated regime). For small φ (i.e., for a small-timescale ratio tmk / tmm̃ ), the trajectory is attracted by the K-vertex and, conversely,
by the M̃ -vertex for large φ. The correct topology of the parametric space—a quadrilateral, not a triangle—was recognized by D.I.
Garagash (personal communication, 2002). (b) Plot of fracture radius γ = R / Lmk (solid lines) versus time τ = t / tmk for φ = 10−4 , 10−2 ,
1, 102 , and 104 . The dashed blue lines are the K̃ -vertex solution for the corresponding values of φ. The solution along the MK-edge
(dashed green line) is also included.

radius R evolves at these vertices according to R ∼ t α , where the exponent α depends on the
regime of propagation: α = 4 / 9 (M), 2/5 (K), and 1/4 ( M̃ and K̃ ).
In view of the dependence of parameters M, K, C, and S on time (see Table 1), it is evident
that the M-vertex corresponds to the small-time asymptote (as Km ∼ t 1/9 and Cm ∼ t 7/18 ) and the
K̃ -vertex to the large-time asymptote (as Mk̃ ∼ t −1/4 and Sk̃ ∼ t −3/8 ). The two other vertices are
intermediate-time asymptotes as Km̃ ∼ t 1/16 and Sm̃ ∼ t −7/16 at the M̃ -vertex and Mk ∼ t −2/5 and
Ck ∼ t 3/10 at the K-vertex. The position of the state point in the MK K̃ M̃-space can be conveniently
expressed in terms of the dimensionless times τmk = t / tmk and τmm̃ = t / tmm̃ , which define the
evolution of the solution along the respective MK- and M M̃-edge. (Obviously, history parameters
−5/18 1/9
P1 and P2 can be expressed in terms of these dimensionless times—e.g., Km = Mk = τmk
−8/9 7/18
and Cm = Sm̃ = τmm̃ .) The evolution of the solution in the MK K̃ M̃-space takes place along a
trajectory corresponding to a constant value of the number φ (0 ≤ φ < ∞), which is related to
the ratio of characteristic times and which depends on all of the problem parameters:
 
E 11 μ3 C 4 Q 0 tmk 14/9
φ= = . (45)
K 14 tmm̃
Examples of such trajectories are depicted in Figure 7a.
In summary, the system evolves with time along a φ-trajectory, starting at the M-vertex
(viscosity-storage-dominated regime: Km = 0, Cm = 0) and ending at the K̃ -vertex (toughness-
leak-off-dominated regime: Mk̃ = 0, Sk̃ = 0). For small values of φ (i.e., for small ratio tmk / tmm̃ ),
the trajectory is attracted by the K-vertex; conversely, for large values of φ, the trajectory is at-
tracted by the M̃-vertex. The evolution of the fracture in the MK K̃ M̃-space is linked, in part, to
the multiscale tip asymptote (Garagash et al. 2011), particularly to the transition from the viscos-
ity M M̃-edge to the toughness K K̃ -edge. Along the viscosity edge, the tip opening progressively
transforms from w ∼ s 2/3 at the M-vertex to w ∼ s 5/8 at the M̃-vertex.
Figures 7b and 8 present computational results for the propagation of a penny-shaped
hydraulic fracture along five different trajectories, φ = 10−4 , 10−2 , 100 , 102 , and 104 , widely

332 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

101 1.0
a b
– 4
e ϕ = 10
dg
-e 0.8
MK

10

4
10

4
0.6

10
Ω (0,τ)

ϕ=

ϕ=
η

0
10 0

10
9 0.4

4
1
0.2

10 –1 –7 0
10 10 –5 10 –3 10 –1 101 103 105 107 10 –7 10 –5 10 –3 10 –1 101 103 105 107
τ τ

Figure 8
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Evolution of (a) fracture opening


= w/W mk at the injection point and (b) of fracture efficiency η with time τ = t / tmk for φ = 10−4 ,
10−2 , 1, 102 , and 104 . The opening solution along the MK-edge (dashed green line) is included in panel a. Figure adapted from
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Madyarova (2004).

scattered inside the parametric space MK K̃ M̃ (Madyarova 2004). The results are reported in the
same time-based MK-scaling used in Section 4.3.3. Evolution of the fracture along the K K̃ −edge
was independently computed by Bunger et al. (2005a).
Figure 7b shows the evolution of the dimensionless fracture radius γ = R/Lmk with time
τ = τmk = t / tmk for different values of φ. The solution along the MK-edge has been included for
comparison. This plot confirms that the solution of γ initially follows the MK-edge solution, as
expected. Departure from the MK-edge solution happens faster for larger φ. It is apparent that γ
clearly evolves from the viscosity-storage propagation regime (γ ∼ τ 4/9 ) toward the toughness-
leak-off regime (γ ∼ τ 1/4 ).
The evolution of the fracture opening
= w/W mk at the injection point is illustrated in
Figure 8a. The solution for a fracture propagating along the storage edge (φ = 0) has also been
included to show that the solutions for different values of φ indeed start from the MK-edge solution
and then gradually deviate from it with time. Figure 8b illustrates the evolution of the efficiency η,
defined as the ratio of the fracture volume over the injected fluid volume, for different values of the
trajectory parameter φ. It can be seen that the efficiency drop is delayed in time with decreasing
φ: A 50% efficiency is reached at time τ 102 for φ = 10−4 but at τ 10−4 for φ = 104 , whereas
a 5% efficiency is reached at τ 105 for φ = 10−4 but at τ 100 for φ = 104 . For example, a
50% efficiency is achieved in approximately 5 h for Q 0 = 2 × 10−2 m3 /s, E = 5 GPa, ν = 0.2,
μ = 0.1 Pa·s, K Ic = 1 MPa·m1/2 , and CL 1.3×10−5 m/s1/2 (corresponding to φ = 10−4 ), but is
achieved in approximately 0.02 s for CL 1.3×10−3 m/s1/2 (φ = 104 ).

5.4. Plane-Strain Fracture


The evolution of a plane-strain hydraulic fracture propagating in a permeable rock with Carter’s
leak-off also can be tracked in the MK K̃ M̃-space (Hu & Garagash 2010). For a constant injection
rate Q 0 , the balance between the viscous dissipation and energy rate expended in the creation of
a new fracture surface does not evolve with time; in other words, the dimensionless toughness
(Km or Km̃ ) or the viscosity (Mk or Mk̃ ) is time invariant. Thus, this class of problems has one
timescale, which characterizes the transition between storage edge MK and leak-off edge M̃K̃ .
These two edges are the locus of similarity solutions, and the solution trajectories are parallel to
the viscosity or toughness edge. The solution for arbitrary toughness Km is described by Hu &

www.annualreviews.org • Mechanics of Hydraulic Fractures 333


FL48CH12-Detournay ARI 3 December 2015 9:24

Garagash (2010), along the viscosity edge M M̃ by Adachi & Detournay (2008), and along the
toughness K K̃-edge by Bunger et al. (2005a).

6. CONCLUSIONS
Competing processes control the propagation of fluid-driven fractures. In impermeable rocks, the
regime of propagation is dictated by the nature of the energy dissipation in the crack-tip region,
where there is competition between viscous dissipation and energy expended in the creation of
new fracture surfaces. In permeable rocks, the fracture evolution is additionally governed by com-
petition between fluid storage in the fracture and leak-off in the surrounding rock to balance the
injected fluid. The interplay between these competing processes transpires through the presence
of velocity-dependent length scales in the tip asymptotics and the existence of timescales in the
fracture evolution.
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Scaling provides a framework to classify and organize regimes of hydraulic fractures, to identify
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

the relevant parameters for specific treatments, or to design laboratory experiments that mimic
field conditions. For simple fracture geometries, the timescales can be expressed explicitly in
terms of the problem parameters (fluid viscosity, rock properties, far-field stress, and injection
rate). Recognition of the existence of multiple timescales has led to the identification of similarity
solutions, which are time asymptotics in the solution trajectory. These solutions, which are as-
sociated with a particular tip asymptote, correspond to regimes of fracture propagation that are
dominated by one dissipation mechanism at the tip and one fluid balance mechanism.
With a focus on the mechanics of hydraulic fractures propagating in a uniform infinite medium,
this review has left untouched issues that are critical to field operations. They include (a) the in-
teraction between closely spaced hydraulic fractures that are created either simultaneously or
sequentially (Olson & Dahi Taleghani 2009, Dahi Taleghani 2011, Bunger et al. 2012, Bunger
2013, Lecampion & Desroches 2015); (b) vertical growth of the fracture into layers subjected to
higher in situ stress (Settari & Cleary 1986, Jeffrey & Bunger 2007, Adachi et al. 2010, Peirce
2015); (c) the propagation of near-surface fractures (Zhang et al. 2002, Bunger & Detournay 2005,
Gordeliy & Detournay 2011b, Bunger et al. 2013a); (d ) the crossing of pre-existing discontinu-
ities such as faults or joints, a crucial issue for the stimulation of naturally fractured reservoirs
(Warpinski & Teufel 1987, Zhang et al. 2007, Zhang & Jeffrey 2008, Jeffrey et al. 2009b); and
(e) the modeling of complex hydraulic fractures in naturally fractured formations (Kresse et al.
2012, Weng 2015).

FUTURE ISSUES
1. Fluid flow models need to be developed to account for changes in the rheology of the
fracturing fluid caused by the addition of proppant (a granular material such as sand that
prevents complete closure of the fracture after treatment), as well as for proppant trans-
port and gravitational settling (Dontsov & Peirce 2014, Lecampion & Garagash 2014).
2. Numerical algorithms must be capable of simulating front recession and fracture closure
during the post-injection phase, arising from either shut-in (Q = 0) or flow-back (Q < 0).
3. Models of hydraulic fracturing induced by water injection taking place over months or
years must account for large-scale 3D diffusion, as well as for poroelastic stress changes
caused by pore pressure perturbations (Zhang et al. 2007, Sarris & Papanastasiou 2012,
Carrier & Granet 2012, Kovalyshen & Detournay 2013).

334 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

4. The need to consider more realistic geological situations requires the development of
simulators that rely on domain-based computational algorithms, such as the extended
finite element method (Gordeliy & Peirce 2013, 2015), the phase-field method (Mikelić
et al. 2015), or the lattice model (Damjanac et al. 2013).

DISCLOSURE STATEMENT
The author is not aware of any biases that might be perceived as affecting the objectivity of this
review.

ACKNOWLEDGMENTS
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

This review reflects, to a great extent, contributions from my former students and postdocs, as
well as from colleagues at CSIRO (Australia) and at UBC (Canada). They are too numerous to
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

be identified here individually, but the cited references are testimony of their contributions. My
interest in this topic was inspired by Professor J.R.A. Pearson, who has the talent of asking the right
questions. Funding has been provided over the years by CSIRO, Schlumberger, the Petroleum
Research Fund, and the National Science Foundation.

LITERATURE CITED
Abé H, Keer LM, Mura T. 1979. Theoretical study of hydraulically fractured penny-shaped cracks in hot, dry
rocks. Int. J. Numer. Anal. Methods Geomech. 3:79–96
Abé H, Mura T, Keer LM. 1976. Growth rate of a penny-shaped crack in hydraulic fracturing of rocks.
J. Geophys. Res. Solid Earth 81:5335–40
Abramowitz M, Stegun IA, eds. 1972. Handbook of Mathematical Functions with Formulas, Graphs, and Mathe-
matical Tables. New York: Dover
Adachi JI. 2001. Fluid-driven fracture in permeable rock. PhD Thesis, Univ. Minnesota, Minneapolis
Adachi JI, Detournay E. 2002. Self-similar solution of a plane-strain fracture driven by a power-law fluid. Int.
J. Numer. Anal. Methods Geomech. 26:579–604
Adachi JI, Detournay E. 2008. Plane strain propagation of a hydraulic fracture in a permeable rock. Eng. Fract.
Mech. 75:4666–94
Adachi JI, Detournay E, Peirce AP. 2010. Analysis of the classical pseudo-3D model for hydraulic fracture
with equilibrium height growth across stress barriers. Int. J. Rock Mech. Min. Sci. 47:625–39
Adachi JI, Siebrits E, Peirce AP, Desroches J. 2007. Computer simulation of hydraulic fractures. Int. J. Rock
Mech. Min. Sci. 44:739–57
Advani SH, Lee TS, Lee JK. 1990. Three-dimensional modeling of hydraulic fractures in layered media: finite
element formulations. ASME J. Energy Res. Technol. 112:1–9
Advani SH, Torok JS, Lee JK, Choudhry S. 1987. Explicit time-dependent solutions and numerical evaluations
for penny-shaped hydraulic fracture models. J. Geophys. Res. Solid Earth 92:8049–55
Barenblatt GI. 1956. On the formation of horizontal cracks in hydraulic fracture of an oil-bearing stratum.
Prikl. Mat. Mech. 20:475–86
Batchelor GK. 1967. An Introduction to Fluid Dynamics. Cambridge, UK: Cambridge Univ. Press
Bunger AP. 2006. A photometry method for measuring the opening of fluid-filled fractures. Meas. Sci. Technol.
17:3237–44
Bunger AP. 2013. Analysis of the power input needed to propagate multiple hydraulic fractures. Int. J. Solids
Struct. 50:1538–49
Bunger AP, Cruden AR. 2011. Modeling the growth of laccoliths and large mafic sills: the role of magma body
forces. J. Geophys. Res. Solid Earth 116:B02203

www.annualreviews.org • Mechanics of Hydraulic Fractures 335


FL48CH12-Detournay ARI 3 December 2015 9:24

Bunger AP, Detournay E. 2005. Asymptotic solution for a penny-shaped near-surface hydraulic fracture. Eng.
Fract. Mech. 72:2468–86
Bunger AP, Detournay E. 2007. Early time solution for a penny-shaped hydraulic fracture. J. Eng. Mech.
133:534–40
Bunger AP, Detournay E. 2008. Experimental validation of the tip asymptotics for a fluid-driven crack.
J. Mech. Phys. Solids 56:3101–15
Bunger AP, Detournay E, Garagash DI. 2005a. Toughness-dominated hydraulic fracture with leak-off. Int. J.
Fract. 134:175–90
Bunger AP, Detournay E, Garagash DI, Peirce AP. 2007. Numerical simulation of hydraulic fracturing in the
viscosity-dominated regime. Proc. 2007 SPE Hydraul. Fract. Technol. Conf., SPE 110115. Richardson, TX:
Soc. Petrol. Eng.
Bunger AP, Detournay E, Jeffrey RG. 2005b. Crack tip behavior in near-surface fluid-driven fracture experi-
ments. C. R. Méc. 333:299–304
Bunger AP, Gordeliy E, Detournay E. 2013a. Comparison between laboratory experiments and coupled
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

simulations of saucer-shaped hydraulic fractures in homogeneous brittle-elastic solids. J. Mech. Phys.


Solids 61:1636–54
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Bunger AP, Jeffrey RG, Detournay E. 2005c. Experimental investigation of crack opening asymptotics for
fluid-driven fracture. Int. J. Strength Fract. Complex. 3:139–47
Bunger AP, McLennan J, Jeffrey RG, eds. 2013b. Effective and Sustainable Hydraulic Fracturing. Rijeka, Croat.:
InTech
Bunger AP, Zhang X, Jeffrey RG. 2012. Parameters effecting the interaction among closely-spaced hydraulic
fractures. Soc. Petrol. Eng. J. 17:292–306
Cameron J, Prud’homme R. 1989. Rheology of fracturing fluids. In Recent Advances in Hydraulic Fracturing,
ed. J Gidley, S Holditch, D Nierode, R Veatch Jr., pp. 177–209. Richardson, TX: Soc. Petrol. Eng.
Carbonell RS, Desroches J, Detournay E. 1999. A comparison between a semi-analytical and a numerical
solution of a two-dimensional hydraulic fracture. Int. J. Solids Struct. 36:4869–88
Carrier B, Granet S. 2012. Numerical modeling of hydraulic fracture problem in permeable medium using
cohesive zone model. Eng. Fract. Mech. 79:312–28
Carter E. 1957. Optimum fluid characteristics for fracture extension. In Drilling and Production Practices, ed.
G Howard, C Fast, pp. 261–70. Tulsa, OK: Am. Petrol. Inst.
Clifton RJ. 1989. Three-dimensional fracture-propagation models. In Recent Advances in Hydraulic Fracturing,
ed. JL Gidley, SA Holditch, DE Nierode, RW Veatch Jr., pp. 95–108. Richardson, TX: Soc. Petrol. Eng.
Clifton RJ, Abou-Sayed AS. 1979. On the computation of the three-dimensional geometry of hydraulic frac-
tures. Proc. 1979 SPE Symp. Low Permeability Gas Reserv., SPE 7943. Richardson TX: Soc. Petrol. Eng.
Constien VG, Hawkins GW, Prud’homme RK, Navarrete R. 2000. Performance of fracturing materials. See
Economides & Nolte 2000, ch. 8
Crouch SL, Starfield AM. 1983. Boundary Element Methods in Solid Mechanics. London: Allen & Unwin
Dahi Taleghani A. 2011. Modeling simultaneous growth of multi-branch hydraulic fractures. Proc. 45th US
Rock Mech. Symp., ARMA 11-436. Alexandria, VA: Am. Rock Mech. Assoc.
Damjanac B, Detournay C, Cundall PA, Varun. 2013. Three-dimensional numerical model of hydraulic
fracturing in fractured rock mass. See Bunger et al. 2013b, ch. 41. doi: 10.5772/56313
Desroches J, Detournay E, Lenoach B, Papanastasiou P, Pearson JRA, et al. 1994. The crack tip region in
hydraulic fracturing. Proc. R. Soc. Lond. A 447:39–48
Detournay E, Garagash DI. 2003. The near-tip region of a fluid-driven fracture propagating in a permeable
elastic solid. J. Fluid Mech. 494:1–32
Detournay E, Peirce AP. 2014. On the moving boundary conditions for a hydraulic fracture. Int. J. Eng. Sci.
84:147–55
Detournay E, Peirce AP, Bunger AP. 2007. Viscosity-dominated hydraulic fractures. In Rock Mechanics: Meeting
Society’s Challenges and Demands, ed. E Eberhardt, D Stead, T Morrison, pp. 1649–56. London: Taylor
& Francis
Dontsov EV, Peirce AP. 2014. Slurry flow, gravitational settling and a proppant transport model for hydraulic
fractures. J. Fluid Mech. 760:567–90

336 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

Economides MJ, Nolte KG, eds. 2000. Reservoir Stimulation. New York: Wiley. 3rd ed.
Emerman SH, Turcotte DL, Spence DA. 1986. Transport of magma and hydrothermal solutions by laminar
and turbulent fluid fracture. Phys. Earth Planet. Interior 36:276–84
Garagash DI. 2000. Hydraulic fracture propagation in elastic rock with large toughness. In Rock Around the Rim:
Proc. 4th N. Am. Rock Mech. Symp., ed. J Girard, M Liebman, C Breeds, T Doe, pp. 221–28. Rotterdam:
Balkema
Garagash DI. 2006a. Plane-strain propagation of a fluid-driven fracture during injection and shut-in: asymp-
totics of large toughness. Eng. Fract. Mech. 73:456–81
Garagash DI. 2006b. Propagation of plane-strain hydraulic fracture with a fluid lag: early-time solution. Int.
J. Solids Struct. 43:5811–35
Garagash DI. 2006c. Transient solution for a plane-strain fracture driven by a power-law fluid. Int. J. Numer.
Anal. Methods Geomech. 30:1439–75
Garagash DI, Detournay E. 2000. The tip region of a fluid-driven fracture in an elastic medium. J. Appl. Mech.
67:183–92
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Garagash DI, Detournay E. 2005. Plane-strain propagation of a fluid-driven fracture: small toughness solution.
J. Appl. Mech. 72:916–28
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Garagash DI, Detournay E, Adachi JI. 2011. Multiscale tip asymptotics in hydraulic fracture with leak-off.
J. Fluid Mech. 669:260–97
Geertsma J, de Klerk F. 1969. A rapid method of predicting width and extent of hydraulic induced fractures.
J. Petrol. Technol. 21:1571–81
Gordeliy E, Detournay E. 2011a. Displacement discontinuity method for modeling axisymmetric cracks in an
elastic half-space. Int. J. Solids Struct. 48:2614–29
Gordeliy E, Detournay E. 2011b. A fixed grid algorithm for simulating the propagation of a shallow hydraulic
fracture with a fluid lag. Int. J. Numer. Anal. Methods Geomech. 35:602–29
Gordeliy E, Peirce A. 2013. Coupling schemes for modeling hydraulic fracture propagation using the XFEM.
Comput. Methods Appl. Mech. Eng. 253:305–22
Gordeliy E, Peirce A. 2015. Enrichment strategies and convergence properties of the XFEM for hydraulic
fracture problems. Comput. Methods Appl. Mech. Eng. 283:474–502
Hills DA, Kelly PA, Dai DN, Korsunsky AM. 1996. Solution of Crack Problems: The Distributed Dislocation
Technique. Dordrecht: Kluwer Acad.
Hu J, Garagash DI. 2010. Plane-strain propagation of a fluid-driven crack in a permeable rock with fracture
toughness. J. Eng. Mech. 136:1152–66
Irwin GR. 1957. Analysis of stresses and strains near the end of a crack traversing a plate. J. Appl. Mech.
29:361–64
Jeffrey RG, Bunger AP. 2007. A detailed comparison of experimental and numerical data on hydraulic fracture
height growth through stress contrasts. Proc. 2007 SPE Hydraul. Fract. Technol. Conf., SPE 106030.
Richardson, TX: Soc. Petrol. Eng.
Jeffrey RG, Bunger AP, Lecampion B, Zhang X, Chen ZR, van As A, Allison DP, et al. 2009a. Measuring
hydraulic fracture growth in naturally fractured rock. Proc. SPE Annu. Tech. Conf. Exhib., SPE124919.
Richardson, TX: Soc. Petrol. Eng.
Jeffrey RG, Chen Z, Mills K, Pegg S. 2013. Monitoring and measuring hydraulic fracturing growth dur-
ing preconditioning of a roof rock over a coal longwall panel. See Bunger et al. 2013b, ch. 45. doi:
10.5772/56325
Jeffrey RG, Zhang X, Thiercelin M. 2009b. Hydraulic fracture offsetting in naturally fractured reservoirs: quan-
tifying a long-recognized process. Proc. 2009 SPE Hydraul. Fract. Technol. Conf., SPE 119351. Richardson,
TX: Soc. Petrol. Eng.
Kanninen MF, Popelar CH. 1985. Advanced Fracture Mechanics. New York: Oxford Univ. Press
Khristianovic SA, Zheltov YP. 1955. Formation of vertical fractures by means of highly viscous fluids. Proc.
4th World Petrol. Congr., Rome, Vol. II, pp. 579–86. London: World Petroleum Council
Kovalyshen Y, Detournay E. 2013. Fluid-driven fracture in a poroelastic rock. See Bunger et al. 2013b, ch.
29. doi: 10.5772/56460

www.annualreviews.org • Mechanics of Hydraulic Fractures 337


FL48CH12-Detournay ARI 3 December 2015 9:24

Kresse O, Weng X, Wu R, Gu H. 2012. Numerical modeling of hydraulic fractures interaction in complex


naturally fractured formations. Proc. 46th US Rock Mech./Geomech. Symp., ARMA-292. Washington, DC:
Am. Rock Mech. Assoc.
Lecampion B, Desroches J. 2015. Simultaneous initiation and growth of multiple radial hydraulic fractures
from a horizontal wellbore. J. Mech. Phys. Solids 82:235–58
Lecampion B, Garagash DI. 2014. Confined flow of suspensions modelled by a frictional rheology. J. Fluid
Mech. 759:197–235
Lecampion B, Peirce A, Detournay E, Zhang X, Chen Z, et al. 2013. The impact of the near-tip logic on
the accuracy and convergence rate of hydraulic fracture simulators compared to reference solutions. See
Bunger et al. 2013b, ch. 43. doi: 10.5772/56212
Lenoach B. 1995. The crack tip solution for hydraulic fracturing in a permeable solid. J. Mech. Phys. Solids
43:1025–43
Lhomme T, Detournay E, Jeffrey RG. 2005. Effects of fluid compressibility and borehole radius on the
propagation of a fluid-driven fracture. Int. J. Strength Fract. Complex. 3:149–62
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Lister JR. 1990. Buoyancy-driven fluid fracture: the effects of material toughness and of low-viscosity precur-
sors. J. Fluid Mech. 210:263–80
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Lister JR, Kerr RC. 1991. Fluid-mechanical models of crack propagation and their application to magma
transport in dykes. J. Geophys. Res. Solid Earth 96:10049–77
Madyarova M. 2004. Propagation of a penny-shaped hydraulic fracture in elastic rock. MSc Thesis, Univ. Minnesota,
Minneapolis
Medlin WL, Masse L. 1984. Laboratory experiments in fracture propagation. Soc. Petrol. Eng. J. 24:256–68
Mendelsohn D. 1984. A review of hydraulic fracture modeling—part I: general concepts, 2D models, moti-
vation for 3D modeling. ASME J. Energy Res. Technol. 106:369–76
Mikelić M, Wheeler MF, Wick T. 2015. A phase-field method for propagating fluid-filled fractures coupled
to a surrounding porous medium. SIAM Multiscale Model. Simul. 13:367–98
Murdoch LC. 2002. Mechanical analysis of idealized shallow hydraulic fracture. J. Geotech. Geoenviron.
128:488–95
Nilson RH. 1986. An integral method for predicting hydraulic fracture propagation driven by gases or liquids.
Int. J. Numer. Anal. Methods Geomech. 10:191–211
Nilson RH, Griffiths SK. 1983. Numerical analysis of hydraulically-driven fractures. Comput. Methods Appl.
Mech. Eng. 36:359–70
Olson JE, Dahi Taleghani A. 2009. Modeling simultaneous growth of multiple hydraulic fractures and their
interaction with natural fractures. Proc. 2009 SPE Hydraul. Fract. Technol. Conf., SPE 119739. Richardson,
TX: Soc. Petrol. Eng.
Osher S, Sethian JA. 1988. Fronts propagating with curvature-dependent speed: algorithms based on
Hamilton-Jacobi formulations. J. Comput. Phys. 79:12–49
Peirce AP. 2010. A Hermite cubic collocation scheme for plane strain hydraulic fractures. Comput. Methods
Appl. Mech. Eng. 199:1949–62
Peirce AP. 2015. Modeling multi-scale processes in hydraulic fracture propagation using the implicit level
set algorithm. Comput. Methods Appl. Mech. Eng. 283:881–908
Peirce AP, Detournay E. 2008. An implicit level set method for modeling hydraulically driven fractures.
Comput. Methods Appl. Mech. Eng. 197:2858–85
Pollard DD, Hozlhausen G. 1979. On the mechanical interaction between a fluid-filled fracture and the Earth’s
surface. Tectonophysics 53:27–57
Rice JR. 1968. Mathematical analysis in the mechanics of fracture. In Fracture: An Advanced Treatise, ed.
H Liebowitz, pp. 191–311. New York: Academic
Rice JR. 1972. Some remarks on elastic crack-tip stress fields. Int. J. Solids Struct. 8:751–58
Roper SM, Lister JR. 2005. Buoyancy-driven crack propagation from an over-pressured source. J. Fluid Mech.
536:79–98
Roper SM, Lister JR. 2007. Buoyancy-driven crack propagation: the limit of large fracture toughness. J. Fluid
Mech. 580:359–80
Rubin AM. 1995. Propagation of magma-filled cracks. Annu. Rev. Earth Planet. Sci. 23:287–336

338 Detournay
FL48CH12-Detournay ARI 3 December 2015 9:24

Sarris E, Papanastasiou P. 2012. Modeling of hydraulic fracturing in a poroelastic cohesive formation. Int. J.
Geomech. 12:160–67
Savitski A, Detournay E. 2002. Propagation of a fluid-driven penny-shaped fracture in an impermeable rock:
asymptotic solutions. Int. J. Solids Struct. 39:6311–37
Sethian JA. 1999. Level Set Methods and Fast Marching Methods: Evolving Interfaces in Computational Geometry,
Fluid Mechanics, Computer Vision, and Materials Science. Cambridge, UK: Cambridge Univ. Press
Settari A. 1985. A new general model of fluid loss in hydraulic fracturing. Soc. Petrol. Eng. J. 25:491–501
Settari A, Cleary MP. 1986. Development and testing of a pseudo-three-dimensional model of hydraulic
fracture geometry. SPE Product. Eng. 1:449–66
Shah KR, Carter BJ, Ingraffea AR. 1997. Hydraulic fracturing simulation in parallel computing environments.
Int. J. Rock Mech. Min. Sci. 34:474
Sneddon IN. 1946. The distribution of stress in the neighbourhood of a crack in an elastic solid. Proc. R. Soc.
Lond. A 187:229–60
Sousa JLS, Carter BJ, Ingraffea AR. 1993. Numerical simulation of 3D hydraulic fracture using Newtonian
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

and power-law fluids. Int. J. Rock Mech. Min. Sci. 30:1265–71


Spence DA, Sharp PW. 1985. Self-similar solution for elastohydrodynamic cavity flow. Proc. R. Soc. Lond. A
Access provided by 165.22.104.127 on 04/16/20. For personal use only.

400:289–313
Spence DA, Sharp PW, Turcotte DL. 1987. Buoyancy-driven crack propagation: a mechanism for magma
migration. J. Fluid Mech. 174:135–53
Tsai VC, Rice JR. 2010. A model for turbulent hydraulic fracture and application to crack propagation at
glacier beds. J. Geophys. Res. 115:F03007
van As A, Jeffrey RG. 2002. Hydraulic fracture growth in naturally fractured rock: mine through mapping
and analysis. In Mining and Tunn. Innov. Oppor.: Proc. 5th N. Am. Rock Mech. Symp. 17th Tunn. Assoc. Can.
Conf.: NARMS-TAC 2002, ed. R Hammah, W Bawden, J Curran, M Telesnicki, pp. 1461–69. Toronto:
Univ. Toronto Press
van Dam DB, de Pater CJ, Romijn R. 1999. Reopening of hydraulic fractures in laboratory experiments. Proc.
9th Int. Congr. Rock Mech., Vol. 2, pp. 791–94. Rotterdam: Balkema
Vandamme L, Curran JH. 1989. A three-dimensional hydraulic fracturing simulator. Int. J. Numer. Methods
Eng. 28:909–27
Voller VR. 2009. Basic Control Volume Finite Element Methods for Fluids and Solids. Singapore: World Sci.
Warpinski NR, Teufel LW. 1987. Influence of geologic discontinuities on hydraulic fracture propagation.
J. Petrol. Technol. 39:209–20
Weng X. 2015. Modeling of complex hydraulic fractures in naturally fractured formation. J. Unconv. Oil Gas
Resour. 9:114–35
Williams ML. 1952. Stress singularities resulting from various boundary conditions in angular corners of
plates in extension. J. Appl. Mech. 19:526–28
Zhang X, Detournay E, Jeffrey RG. 2002. Propagation of a penny-shaped hydraulic fracture parallel to the
free-surface of an elastic half-space. Int. J. Fract. 115:125–58
Zhang X, Jeffrey RG. 2008. Reinitiation or termination of fluid-driven fractures at frictional bedding interfaces.
J. Geophys. Res. 113:B08416
Zhang X, Jeffrey RG, Thiercelin M. 2007. Deflection and propagation of fluid-driven fractures at frictional
bedding interfaces: a numerical investigation. J. Struct. Geol. 29:396–410

www.annualreviews.org • Mechanics of Hydraulic Fractures 339


FL48-FrontMatter ARI 3 December 2015 9:31

Annual Review of
Fluid Mechanics

Volume 48, 2016 Contents


Biomimetic Survival Hydrodynamics and Flow Sensing
Michael S. Triantafyllou, Gabriel D. Weymouth, and Jianmin Miao p p p p p p p p p p p p p p p p p p p p p p 1
Motion and Deformation of Elastic Capsules and Vesicles in Flow
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Dominique Barthès-Biesel p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p25


Access provided by 165.22.104.127 on 04/16/20. For personal use only.

High–Reynolds Number Taylor-Couette Turbulence


Siegfried Grossmann, Detlef Lohse, and Chao Sun p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p53
Shear Banding of Complex Fluids
Thibaut Divoux, Marc A. Fardin, Sebastien Manneville, and Sandra Lerouge p p p p p p p p p p p81
Bacterial Hydrodynamics
Eric Lauga p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 105
Quadrant Analysis in Turbulence Research: History and Evolution
James M. Wallace p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 131
Modeling of Fine-Particle Formation in Turbulent Flames
Venkat Raman and Rodney O. Fox p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 159
Seismic Sounding of Convection in the Sun
Shravan Hanasoge, Laurent Gizon, and Katepalli R. Sreenivasan p p p p p p p p p p p p p p p p p p p p p p 191
Cerebrospinal Fluid Mechanics and Its Coupling to Cerebrovascular
Dynamics
Andreas A. Linninger, Kevin Tangen, Chih-Yang Hsu, and David Frim p p p p p p p p p p p p p p 219
Fluid Mechanics of Heart Valves and Their Replacements
Fotis Sotiropoulos, Trung Bao Le, and Anvar Gilmanov p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 259
Droplets and Bubbles in Microfluidic Devices
Shelley Lynn Anna p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 285
Mechanics of Hydraulic Fractures
Emmanuel Detournay p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 311
A Normal Mode Perspective of Intrinsic Ocean-Climate Variability
Henk Dijkstra p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 341
Drop Impact on a Solid Surface
C. Josserand and S.T. Thoroddsen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 365

vi
FL48-FrontMatter ARI 3 December 2015 9:31

Contrail Modeling and Simulation


Roberto Paoli and Karim Shariff p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 393
Modeling Nonequilibrium Gas Flow Based on Moment Equations
Manuel Torrilhon p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 429
The Fluid Mechanics of Pyroclastic Density Currents
Josef Dufek p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 459
The Dynamics of Microtubule/Motor-Protein Assemblies
in Biology and Physics
Michael J. Shelley p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 487
Dynamics and Instabilities of Vortex Pairs
Annu. Rev. Fluid Mech. 2016.48:311-339. Downloaded from www.annualreviews.org

Thomas Leweke, Stéphane Le Dizès, and Charles H.K. Williamson p p p p p p p p p p p p p p p p p p p p p 507


Access provided by 165.22.104.127 on 04/16/20. For personal use only.

Indexes

Cumulative Index of Contributing Authors, Volumes 1–48 p p p p p p p p p p p p p p p p p p p p p p p p p p p p 543


Cumulative Index of Article Titles, Volumes 1–48 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 553

Errata

An online log of corrections to Annual Review of Fluid Mechanics articles may be


found at http://www.annualreviews.org/errata/fluid

Contents vii

Potrebbero piacerti anche