You are on page 1of 28

Received: 16 August 2021 Revised: 30 March 2022 Accepted: 2 May 2022

DOI: 10.1002/nag.3383

RESEARCH ARTICLE

An enhanced J-integral for hydraulic fracture mechanics


Edoardo Pezzulli1 Morteza Nejati1 Saeed Salimzadeh2
Stephan K. Matthäi3 Thomas Driesner1

1 Department of Earth Sciences, ETH


Zurich, Switzerland Abstract
2 Commonwealth Scientific and Industrial This article revisits the formulation of the J-integral in the context of hydraulic
Research Organisation (CSIRO), Australia fracture mechanics. We demonstrate that the use of the classical J-integral in
3 Department of Infrastructure finite element models overestimates the length of hydraulic fractures in the
Engineering, The University of
Melbourne, Victoria, Australia viscosity-dominated regime of propagation. A finite element analysis shows
that the inaccurate numerical solution for fluid pressure is responsible for
Correspondence
the loss in accuracy of the J-integral. With this understanding, two novel
Edoardo Pezzulli, Department of Earth
Sciences, ETH Zurich, Switzerland. contributions are presented. The first contribution consists of two variations
𝐴
Email: edoardo.pezzulli@erdw.ethz.ch of the J-integral, termed the 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 -integral, that demonstrate an
enhanced ability to predict viscosity-dominated propagation. In particular, such
Funding information
Eidgenossische Technische Hochschule 𝐽𝐻𝐹𝑀 -integrals accurately extract stress intensity factors in both viscosity and
Zurich; Schweizerischer Nationalfonds toughness-dominated regimes of propagation. The second contribution consists
zur Förderung der Wissenschaftlichen
Forschung, Grant/Award Number:
of a methodology to extract the propagation velocity from the energy release
200020_204499 rate applicable throughout the toughness-viscous propagation regimes. Both
techniques are combined to form an implicit front-tracking 𝐽𝐻𝐹𝑀 -algorithm
capable of quickly converging on the location of the fracture front independently
to the toughness-viscous regime of propagation. The 𝐽𝐻𝐹𝑀 -algorithm represents
an energy-based alternative to the aperture-based methods frequently used with
the Implicit Level Set Algorithm to simulate hydraulic fracturing. Simulations
conducted at various resolutions of the fracture suggest that the new approach is
suitable for hydro-mechanical finite element simulations at the reservoir scale.

KEYWORDS
energy release rate, finite element, hydraulic fracturing, hydraulic fracture mechanics,
J-integral

1 INTRODUCTION

The propagation of fractures driven by a viscous fluid plays a pivotal role within both geological and geo-engineering con-
texts. The geological relevance of such hydraulic fracturing processes can be seen from the extensive and dynamic cycling
of fluid and melt that occurs at the crustal scale.1–3 and also by the creation of hydrothermal ore-deposits and geothermal
systems that occur at the reservoir scale.4–6 On the othe hand, the engineering of hydraulic fractures is fundamental to the
optimisation of enhanced geothermal systems7,8 and their associated seismicity,9,10 while preventing hydraulic fracturing

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the
original work is properly cited.
© 2022 The Authors. International Journal for Numerical and Analytical Methods in Geomechanics published by John Wiley & Sons Ltd.

Int J Numer Anal Methods Geomech. 2022;46:2163–2190. wileyonlinelibrary.com/journal/nag 2163


2164 PEZZULLI et al.

HIGHLIGHTS
∙ The J-integral gives inaccurate results in the viscous-dominated propagation regime.
∙ A modified J-integral for hydraulic fracturing is proposed that eliminates the error.
∙ The location of the fracture tip is predicted from the energy release rate concept.
∙ An efficient energy-based propagation algorithm for hydraulic fracturing is detailed.

is critical to the integrity of the cap-rock for the secure sequestration of carbon dioxide.11 In such applications, observation
and monitoring is often limited to incomplete and/or indirect methods such as a posteriori imaging of borehole walls or
induced seismicity recordings. Consequently, a main challenge regards quantifying the effect hydraulic fracturing has in
redistributing stress, heat and/or dissolved minerals.
Numerical simulations have the potential to illuminate the effects of hydraulic fractures in subsurface processes. For
such simulations, hydraulic fracture mechanics (HFM) has become the preferred framework for assessing the propagation
of hydraulic fractures. Linear hydraulic fracture mechanics (used synonymously with HFM herein), couples linear elastic
fracture mechanics (LEFM) and the lubrication equations to derive solutions for the displacement, stress and fluid pres-
sure profiles relevant to a propagating hydraulic fracture (see 12 for a summary of analytical and semi-analytical solutions).
As a result, the mechanics of hydraulic fracturing has been understood as the competition between two end-member prop-
agation regimes: in the viscous regime the majority of the energy during propagation is dissipated via driving the viscous
fluid while in the toughness regime the energy dissipation is governed by the creation of new fracture surfaces.13,14 Current
analytical solutions are informative, but limited to describing propagation for such end-member regimes, under simplified
geometries and fluid/rock properties.15,16 In practice, hydraulic fracturing can occur in a mixed and varying propagation
regime, and in complicated geometries with varying fluid/rock properties, all of which require accurate and efficient
numerical models to provide valuable insights.
An important approach in numerically simulating the propagation of a hydraulic fracture has involved the use of tip
asymptotes. These are analytical solutions valid at the tip of the propagating fracture that are then used to inform numer-
ical models on the onset of propagation and the subsequent propagation increments. As it became well-established that
such solutions depend on the propagation regime of the fracture,12–14,17 one of the key challenges for numerical simula-
tions in HFM was to derive a propagation strategy that can equally well be applied to the viscous and toughness regime and
the transition between them.18 Among the possible solutions to this problem,19–22 the aperture asymptote of Dontsov and
Peirce23 constitutes the latest approach in predicting propagation throughout the toughness-viscous-leak-off regimes,24
providing a general solution adopted by subsequent studies.25–32
A salient result that has emerged from HFM is that the tip asymptotes incorporate information on the propagation veloc-
ity of the fracture.17,33 This is a consequence of the conservation of mass of a viscous fluid, which limits the propagation
of the fracture to the fluid’s capacity to fill the additional space. Tip asymptotes that incorporate this viscous effect benefit
from a more accurate computation of stress intensity factors in the viscous-dominated regime; a key parameter whose
computation is necessary to simulate propagation and is often a computational bottleneck in numerical models.12,18,19
Furthermore, such asymptotes allow for the estimation of the propagation velocity, thus enabling an efficient iteration on
the location of the fracture front19,20,24,28,31,34–36 ; the implicit level set algorithm is a notable example.34,37 Using aperture
asymptotes represents the dominant methodology in HFM that exploits such computational and algorithmic advantages.
LEFM can provide a useful analogue to inspire alternative approaches to the current aperture-based methods in HFM.
In particular, the use of the aperture asymptote in computational HFM can be considered as the hydraulic analogue of the
displacement correlation method in LEFM.38–41 Such methods have the disadvantage of being susceptible to local errors in
the numerical solution of the aperture at the fracture tip. Adopting energy-based theories of failure in LEFM avoids these
disadvantages and, instead, achieves significantly higher levels of accuracy and lower resolution requirements.38,42–44 The
quintessential example of such efficient energy-based approaches is the 𝐽-integral, which estimates the energy release rate
of an extending self-planar crack.42,45–52 Unlike the displacement correlation method, the 𝐽-integral is more acceptably
extensible to three -dimensional problems, and constitutes an important method in assessing failure in both non-linear
and time-dependent materials.41
The 𝐽-integral has been applied in the hydraulic fracturing context,53–60 along with alternative energy-based
approaches,61,62 however it has remained elusive whether the aforementioned advantages can be exploited. In
PEZZULLI et al. 2165

particular, these include whether the 𝐽-integral can estimate the stress intensity factor independently to the toughness-
viscous propagation regime, and importantly, whether the 𝐽-integral can also be used to extract the propagation velocity
of the fracture. This present study explores these questions and demonstrates that both can be accurately and efficiently
solved with a 𝐽-integral reformulation. Such a result facilitates an energy-based analysis of failure of hydraulic fracturing
and broadens the current modus operandi of computational HFM.
In this paper, a numerical approach is employed to answer the questions postulated above. Section 2 formulates the
hydraulic fracturing problem, along with the current 𝐽-integral methodology. Section 3 details the reformulations of the
𝐽-integral to 𝐽𝐻𝐹𝑀 -integrals that are suitable for computational HFM. Section 4 describes the finite element model used to
discretise both the elasticity and lubrication equations in order to simulate the Kristianovich–Geertsma–de Klerk (KGD)
problem in hydraulic fracturing. Such a finite element model forms the testing bed for the analysis of the accuracy and
efficiency of the integrals. Section 5 addresses the question concerning the ability of the 𝐽-integral and its HFM variations
in extracting stress intensity factors in the viscous regime of propagation. Finally, Section 6 introduces a methodology
that uses the 𝐽𝐻𝐹𝑀 -integrals to extract the propagation velocity and converge on the fracture front, coined the 𝐽𝐻𝐹𝑀 -
algorithm. A significant implication of the developments presented in this paper is that the 𝐽𝐻𝐹𝑀 -integrals and the 𝐽𝐻𝐹𝑀 -
algorithm can accurately extract the stress intensity factor and the propagation velocity throughout the toughness-viscous
propagation regime, even with a coarse discretisation of the fracture. Such capabilities have the potential to be utilised
within efficient front-tracking schemes like the implicit level set algorithm,34 and open up new possibilities for energy-
based analysis of failure in HFM.

2 THE HYDRAULIC FRACTURING PROBLEM AND THE CLASSICAL J-INTEGRAL

Solving the hydraulic fracturing problem involves obtaining the stresses and displacements in the rock, the pressure and
velocities of the fluid, and the geometry of the evolving fracture. In this work, the focus is on the application of the 𝐽-
integral to predict fracturing within rock that deforms in plane-strain conditions according to the theory of linear elasticity
under infinitesimal strains. The inertia in the rock and the fluid are neglected, and the rock is idealised as impermeable
with isotropic and homogeneous mechanical properties, while the fluid is considered incompressible, with a constant
Newtonian viscosity. No leak-off is occurring from the fracture to the rock, and the fluid lag at the fracture tip is neglected.
The effects of gravity are not included. The assumption of zero-leak off, and the omission of gravity can be relaxed within
the framework proposed in Section 3, but is not subject of this study. Finally, the assumption of a negligible fluid lag is
appropriate for the deep applications mentioned in Section 1, since the size of the lag zone decreases proportionally with
the cube of the confining stress.63,64
With the aforementioned assumptions, hydraulic fracturing can be completely characterised given the fluid pressure 𝑝𝑓
in the fracture, the vector displacements in the rock 𝐮, and the geometry of the fracture Γ𝑐 . A geometrical conceptualisation
of the problem is depicted in Figure 1. Displacements 𝐮 are defined on a rock body Ω that is encapsulated by the surface ΓΩ
and subject to either traction 𝐭Ω or displacement 𝐮Ω boundary conditions acting on the external surface ΓΩ . The fracture
surfaces Γ+ −
𝑐 , Γ𝑐 are delineated by displacements 𝐮+ , 𝐮− , respectively, allowing for a discontinuous displacement field upon
opening of the fracture. The fracture has unit normal 𝒏 to the one-dimensional curve Γ𝑐 that sits on the fractures mid-line,
+
and points by convention from Γ− 𝑐 to Γ𝑐 . The fluid is conceptualised to lie on Γ𝑐 , while exerting its pressure on Γ− and Γ+ .
The solutions for the field variables (𝐮, 𝑝𝑓 ) must be obtained given a geometry of the fracture Γ𝑐 . However, the frac-
ture geometry is itself unknown and must be found. Therefore, once field variables are obtained for a certain fracture
configuration, a post-processing step is required to assess whether the field variables predict stresses above the materials
fracture toughness, and/or propagation velocities different to the simulated velocity implied by the current choice of the
fracture geometry Γ𝑐 . The former can be checked by ensuring the fracture toughness 𝐾𝐼𝑐 is greater or equal to the mode
𝐼 stress intensity factor 𝐾𝐼 at that fracture tip that is predicted from the field variables (𝐮, 𝑝𝑓 ). When 𝐾𝐼 > 𝐾𝐼𝑐 , the mate-
rial sustains levels of stress higher than it can theoretically resist, thus the fracture must propagate. On the other hand,
predicting the propagation velocity is a feature particular to HFM. For a non-zero fluid viscosity, the propagation velocity
is constrained by the viscous fluid’s ability to fill the newly created space upon fracturing, and can be estimated from the
field variables as a result. Such a propagation velocity should be equal to the propagation velocity implied by the fracture
geometry used at the current time step Γ𝑐 . The complete solution at a certain moment in time is illustrated in Figure 2
and consists of the field variables whose solution was obtained for a fracture configuration Γ𝑐 that predicts stress intensity
factors 𝐾𝐼 ≲ 𝐾𝐼𝑐 , and (possibly) a propagation velocity identical to the simulated velocity of the fracture. The 𝐽-integral is
a technique that uses the field variables (𝐮, 𝑝𝑓 ) at the fracture tip to calculate the stress intensity factor 𝐾𝐼 for comparison
2166 PEZZULLI et al.

F I G U R E 1 Mathematical conceptualisation of a fracture Γ𝑐 contained within a domain Ω. The fracture is a one-dimensional curve
hosting discontinuous displacements 𝐮+ , 𝐮−

F I G U R E 2 Illustration of the complete solution to the hydraulic fracturing problem for a particular moment in time. The solution to the
field variables 𝑝𝑓 and 𝒖 (only 𝑢𝑦 illustrated) must be found for a fracture geometry Γ𝑐 which give stress intensity factors 𝐾𝐼 compatible with
the fracture toughness 𝐾𝐼𝑐 , and a propagation velocity 𝑉 that is compatible with the admissible velocity of the fluid at the fracture tip

with 𝐾𝐼𝑐 . The variations of the 𝐽-integral introduced in Section 3 offer the same capabilities, with the addition of being
able to calculate, and check the compatibility of, the propagation velocity of the fracture.
For the remainder of this section, the governing equations for the field variables are detailed, along with the propagation
regimes which characterise the nature of the solution, and the standard procedure used by the 𝐽-integral to extract stress
intensity factors in hydraulic fracturing.
PEZZULLI et al. 2167

2.1 Governing equations

In view of the assumptions discussed in the preceding section, the displacements in the rock are constrained by the static
force equilibrium equations in the absence of body forces

∇ ⋅ 𝝈 = 0, (1a)

where 𝝈 represents the stress in the rock and relates to displacements 𝒖 via the framework of linear elasticity under
infinitesimal strains.65 The fracture’s aperture is described by 𝑤 = (𝒖+ − 𝒖− ) ⋅ 𝒏, where the quantities 𝒖− and 𝒖+ repre-
+
sent discontinuous displacement vectors lying on Γ𝑐 which delineate the location of the bottom (Γ− 𝑐 ) and top (Γ𝑐 ) fracture
surfaces. The aperture 𝑤 is coupled to the fluid pressure 𝑝𝑓 via the Reynolds-lubrication equation
( )
𝜕𝑤 𝜕 𝑤 3 𝜕𝑝𝑓
− = 𝑄0 𝛿(𝑠), (1b)
𝜕𝑡 𝜕𝑠 𝜇′ 𝜕𝑠

where 𝑠 is a coordinate in the direction aligned along the fracture Γ𝑐 . The Newtonian dynamic viscosity 𝜇 is reformulated
in terms of the lumped parameter 𝜇′ = 12𝜇. The Dirac delta function 𝛿 is used to denote the location of the constant point
volume source 𝑄0 entering the fracture.

2.2 Boundary conditions in the absence of a fluid lag

In the absence of a fluid-lag, the fluid pressure at the fracture tip is unknown, requiring instead the fluid flux 𝑞𝑓

𝑤 3 𝜕𝑝𝑓
𝑞𝑓 = − (2)
𝜇′ 𝜕𝑠

to vanish at the tip. Such a condition is necessary to ensure finite average fluid velocities 𝑉𝑓 = 𝑞𝑓 ∕𝑤 as the aperture also
vanishes (𝑤 = 0) at the fracture tip.66 An important characteristic to bear in mind is that the fluid-front and the fracture-
front coalesce under the assumption of zero fluid-lag. Therefore, the propagation velocity of the fracture 𝑉 is identical to
the average velocity of the fluid 𝑉 = 𝑉𝑓 , which significantly simplifies the asymptotic solutions and the extraction of the
propagation velocity 𝑉 from the field variables.

2.3 Propagation regimes

Hydraulic fractures can be classified to propagate according to an array of propagation regimes.14 The most fundamental
of these are the toughness and viscous propagation regimes, where the dimensionless toughness 

( )1∕4
𝐾 ′4
= , (3)
𝐸 ′3 𝜇′ 𝑄

quantifies in which regime a plane-strain fracture is propagating. The lumped parameters are 𝐾 ′ = 32∕𝜋𝐾𝐼 and 𝐸 ′ =
𝐸∕(1 − 𝜈2 ), while 𝑄 is the fluid flux entering the fracture. Values of  ≲ 0.41 implies viscosity-dominated propagation,
while  ≳ 4.13 implies toughness-dominated propagation.
The solution for the stress and displacement at the fracture tip is described by a toughness asymptote close to the fracture
tip (as described by LEFM67,68 ) that transitions to a viscous asymptote further away from the tip.17,33,63 The transition
lengthscale

𝐾 ′6
𝓁𝑘𝑚 = , (4)
𝐸 ′4 𝜇′2 𝑉 2
2168 PEZZULLI et al.

(A) (B) (C)

F I G U R E 3 (A) Schematic showing the coordinate system at the crack tip relevant to the calculation of the 𝐽-integral, (B) Surface plot of
the plateau 𝑞𝐹𝑀 function over two sets of element rings surrounding the fracture tip. (C) The plateau 𝑞̂ 𝐹𝑀 function, which is a one
dimensional slice of 𝑞𝐹𝑀 at the crack faces reformulated as a function of the inward distance from the fracture tip 𝑥̂

where 𝑉 is the propagation velocity, describes at what physical distance inwards from the tip the solution transitions from
the toughness asymptote to the viscous one.12,14,17 Which asymptote is captured by the numerical solution depends on the
resolution at the fracture tip ℎ in comparison to the transition lengthscale 𝓁𝑘𝑚 .
Propagation in the toughness-dominated regime practically ensures that a reasonable resolution of the fracture tip
captures the toughness asymptote. Consequently, the problem bears a close resemblance to LEFM, where the 𝐽-integral
performs to a high level of accuracy.38,41,43,57 In the viscosity-dominated regime, it becomes impractical to capture the
toughness-asymptote by employing mesh resolutions ℎ << 𝓁𝑘𝑚 . Whether the 𝐽-integral retains its accuracy in such a
scenario is investigated. The remainder of this chapter details the current LEFM methodology when applying the 𝐽-integral
in the hydraulic fracturing context.

2.4 Classical J-integral in LEFM

The 𝐽-integral is capable of extracting the stress intensity factor 𝐾𝐼 from the field variables 𝑝𝑓 , 𝒖 that lie within a domain
that surrounds the fracture tip (as illustrated in Figure 2). The 𝐽-integral is best described by adopting a coordinate system
(𝑥1 , 𝑥2 ) with origin at the moving crack tip as shown in Figure 3A. A contour Γ circles around the tip from the lower crack
face to the upper one with outward pointing normal 𝐦. The following integral over Γ
( )
𝜕𝑢𝑗
𝐽Γ = 𝑊 𝛿1𝑖 − 𝜎𝑖𝑗 𝑚𝑖 𝑑Γ, (5)
∫Γ 𝜕𝑥1

is related to the rate of mechanical energy flux entering the crack tip in the absence of body forces.42,45,46 The quantity 𝑊 =
𝜀
∫0 𝑖𝑗 𝜎𝑖𝑗 𝑑𝜀𝑖𝑗 denotes the strain energy density, where 𝜀𝑖𝑗 and 𝛿𝑖𝑗 are the strain and Kronecker delta at the 𝑖, 𝑗 component
which corresponds to the 𝑥, 𝑦 dimension. The summation convention is implied throughout. As the contour Γ shrinks
into the fracture tip Γ → 0, the integral in Equation 5 converges to the energy release rate 𝐺, and consequently the stress
intensity factor 𝐾𝐼 by42

𝐾𝐼2
𝐽 = lim 𝐽Γ = 𝐺 = , (6)
Γ→0 𝐸′

for a linear elastic medium with finite (non-zero) toughness. The relationship with the stress intensity factor in Equation 6
is guaranteed by the existence of the toughness asymptotes at the fracture tip,41,42 which is always the case for a hydraulic
fracture with non-zero fracture toughness.63
Once 𝐽 is evaluated numerically, the stress intensity factor 𝐾𝐼 can be obtained using Equation 6. However, evaluating 𝐽
over a vanishingly small contour Γ is cumbersome to carry out numerically. Instead, the integral’s path-independence is
used to evaluate 𝐽 along the remote contour Γ0
( )
𝜕𝑢𝑗 𝜕𝑢
𝐽= 𝑊 𝛿1𝑖 − 𝜎𝑖𝑗 𝑚𝑖 𝑑Γ + 𝑝𝑓 2 𝑚2 𝑑Γ, (7)
∫Γ 𝜕𝑥1 ∫Γ +Γ 𝜕𝑥1
0 + −
PEZZULLI et al. 2169

along with a contribution over the fracture surfaces Γ+ and Γ− , which lie on the 𝑥1 axes for infinitely sharp cracks and have
normal components 𝑚2 = −1 on Γ+ , and 𝑚2 = 1 on Γ− (they are separated only for illustration purposes in Figure 3A).
The integration over contour Γ0 assumes the relevant outward pointing normal 𝒎. The normal stress at the fracture
surfaces Γ+ and Γ− is 𝜎22 = −𝑝𝑓 , while the shear stresses are significantly smaller at the fracture surfaces and have been
neglected.69 The 𝐽-integral shown in Equation 7 is said to be path-independent in the sense that any contour Γ0 can be
chosen, given the contributions over Γ+ and Γ− are correctly accounted for.
This path-independence property of the 𝐽-integral underpins its accuracy in evaluating the stress intensity factor in
‘dry’ fractures, even at coarse resolutions.38 Such a property is a consequence of quantifying the rate of mechanical energy
flux entering the fracture tip,45,46 which does not rely on the existence of any particular solution of the displacement and
stress profile, and therefore is theoretically independent to the transition length-scale 𝓁𝑘𝑚 .42,45,46 The numerical results
in Sections 5 and 7 further evidence such a feature.
Typically, the 𝐽-integral is expressed in its domain equivalent form, since integration over an area is more compatible
with the finite element method. The domain-equivalent integral of Equation 7 is
( ) 𝐹𝑀
𝜕𝑢𝑗 𝜕𝑞 𝜕𝑢
𝐽= 𝜎𝑖𝑗 − 𝑊𝛿1𝑖 𝑑𝐴 + 𝑝𝑓 2 𝑚2 𝑞𝐹𝑀 𝑑Γ, (8)
∫𝐴 ∗ 𝜕𝑥1 𝜕𝑥𝑖 ∫Γ +Γ 𝜕𝑥1
+ −
⏟⎴⎴⎴⎴⎴⎴⎴⎴⎴⎴⏟⎴⎴⎴⎴⎴⎴⎴⎴⎴⎴⏟ ⏟⎴⎴⎴⎴⎴⎴⎴⏟⎴⎴⎴⎴⎴⎴⎴⏟
𝐽𝑟 𝐽𝑓

where 𝐽𝑟 and 𝐽𝑓 are related to the mechanical energy input via the stress in the rock, and the pressure in the fluid, respec-
tively. The quantity 𝐴∗ is the area enclosed by the contours Γ0 + Γ+ + Γ− − Γ, while 𝑞𝐹𝑀 is a typically used scalar function
in the field of fracture mechanics which must be 1 at the fracture tip and 0 on the outer contour Γ0 , which effectively carries
out a weighted average of the integrands.41,48 Typical choices for 𝑞𝐹𝑀 are the pyramid, or plateau functions41 for which
more will be discussed in Section 3. The superscripts in 𝑞𝐹𝑀 are specifically used to highlight the fracture mechanic origins
of the parameter, and therefore avoid confusion with the fluid flux 𝑞𝑓 .
The 𝐽-integral as written in Equation 8 constitutes the standard formulation commonly used to evaluate the energy
release rate in the hydraulic fracturing context. The accuracy of the numerical solution for fluid pressure 𝑝𝑓 and displace-
ment 𝒖 determine the accuracy of 𝐽. There is no guarantee that a numerical model captures 𝑝𝑓 and 𝒖 with the same level
of accuracy, and therefore, it is of interest to explore reformulations of the 𝐽-integral which exchange one field variable
for another. Such a feature will be seen to be advantageous for the hydraulic fracturing problem, and is possible since the
fluid pressure and displacements are coupled via the Reynolds-lubrication Equation 1b; a coupling not available in LEFM.
Incorporating the Reynolds-lubrication equation within the 𝐽-integral is what specialises the integral for use within HFM.

3 A J-INTEGRAL FOR HYDRAULIC FRACTURE MECHANICS

In this section, the Reynolds-lubrication equation is used to reformulate the 𝐽-integral in order to enhance the accuracy
of computed stress intensity factors 𝐾𝐼 from numerical solutions. In particular, a 𝐽𝐻𝐹𝑀 -integral is introduced which relies
only on numerical values for the aperture near the fracture tip, and not fluid pressure.
The derivation of the 𝐽𝐻𝐹𝑀 -integral rests on using the Reynolds-lubrication equation with the coordinate system 𝑥̂ =
𝑉𝑡 − 𝑠 moving with the fracture tip at a constant speed 𝑉.12 An illustration of 𝑥̂ is shown in Figure 3A. Following earlier
works,17,66 a stationary Reynolds-lubrication equation can be derived in terms of 𝑥̂ to be

𝜕𝑝𝑓
𝜇′ 𝑉 = 𝑤 2 , (9)
𝜕 𝑥̂

where the propagation velocity 𝑉 is also equal to the average velocity of the fluid, since the fluid-lag is neglected.
The stationary Reynold’s Equation 9 plays a profound role in hydraulic fracture mechanics, and reflects that the dis-
sipation of energy due to viscous forces constrains the velocity of propagation. The derivation of Equation 9 is concisely
elucidated by Detournay and Peirce.66 Importantly, the existence of Equation 9 is the cause for the velocity-dependence
of the asymptotic solutions for the fracture aperture and fluid pressure at the fracture tip.17 Significantly, while the fluid
pressure is singular at the fracture tip, it can be exchanged for a finite propagation velocity 𝑉 using Equation 9. Such a
feature was exploited by Dontsov and Peirce23 and will also be exploited here.
2170 PEZZULLI et al.

Since the domain 𝐽-integral is of more interest to finite element simulations, the 𝐽-integral in Equation 8 is the starting
point of our reformulation. A derivation of an equivalent contour version of the 𝐽𝐻𝐹𝑀 -integral is given in A.1. The final
form of the domain 𝐽𝐻𝐹𝑀 -integral will be

𝐽𝐻𝐹𝑀 = 𝐽𝑟 + 𝐽𝑓𝐻𝐹𝑀 , (10)

where 𝐽𝑟 is unchanged from Equation 8, and 𝐽𝑓 is reformulated to 𝐽𝑓𝐻𝐹𝑀 . Focusing on 𝐽𝑓 , the integration along the contours
Γ+ and Γ− can be expressed in terms of 𝑥̂ following Song and Rahmann52

𝑥̂ 0
𝜕𝑤 𝐹𝑀
𝐽𝑓 = 𝑝𝑓 ̂
𝑞 𝑑𝑥, (11)
∫0 𝜕 𝑥̂

where 𝑥̂ 0 is the distance from the fracture tip at which the fracture surfaces intersect with Γ0 (see Figure 3A). Integration
by parts gives

𝑥̂ 0
[ ]𝑥=
̂ 𝑥̂ 0 𝜕
𝐽𝑓 = 𝑝𝑓 𝑤𝑞𝐹𝑀 𝑥=0 − 𝑤 (𝑝 𝑞𝐹𝑀 ) 𝑑𝑥̂ . (12)
̂ ∫0 𝜕 𝑥̂ 𝑓

Evaluation of the first term on the right hand side of Equation 12 gives zero since 𝑤 = 0 at the fracture tip 𝑥̂ = 0, and
𝑞𝐹𝑀 = 0 along the outer contour Γ0 , which is the case when 𝑥̂ = 𝑥̂ 0 . Thus 𝐽𝑓 can be simplified to

𝑥̂ 0 𝑥̂ 0 𝜕𝑝𝑓
𝜕𝑞𝐹𝑀
𝐽𝑓 = − 𝑤𝑝𝑓 𝑑𝑥̂ − 𝑞𝐹𝑀 𝑤 ̂
𝑑𝑥, (13)
∫0 𝜕 𝑥̂ ∫0 𝜕 𝑥̂

where two final modifications are needed to eliminate the dependency of the integral on fluid pressures at the fracture tip.
The first important modification regards choosing a plateau 𝑞𝐹𝑀 function41 that eliminates the need to evaluate fluid
pressure at the fracture tip. It will be shown in Section 5, that numerical solutions for fluid pressure at the fracture tip are
prone to error. Thus, the 𝑞𝐹𝑀 function is used to direct the integration to areas around the fracture tip where numerical
values are most accurate,48 since its choice carries little physical significance apart from carrying out a weighted average of
the integrand.41,48 The plateau 𝑞𝐹𝑀 function is chosen to achieve this aim, since it has a constant value of 1 within a certain
radius from the fracture tip, and decreases linearly to zero on the outer contour Γ0 , as shown in Figure 3B. To understand
how such a choice of 𝑞𝐹𝑀 function changes 𝐽𝑓 in Equation 13, it is beneficial to quantify the value of the plateau 𝑞𝐹𝑀
along the fracture surfaces Γ+ and Γ− . Along such contours on the fracture surface, we have 𝑥2 = 0 and 𝑥1 = −𝑥, ̂ thus the
plateau 𝑞𝐹𝑀 function in Equation 13 can be expressed as 𝑞̂ 𝐹𝑀 = 𝑞𝐹𝑀 (𝑥1 (𝑥), ̂ 𝑥2 = 0), which has a form


⎪1 , if 𝑥̂ < 𝑥̂ 𝑞 ,


𝑞̂ 𝐹𝑀 = ⎨ 𝑥̂ 0 − 𝑥̂ , if 𝑥̂ 𝑞 < 𝑥̂ < 𝑥̂ 0 , (14)
⎪ 𝑥̂ 0 − 𝑥̂ 𝑞

⎪0 , otherwise ,

and is plotted in Figure 3C. The point 𝑥̂ 𝑞 is defined as the distance from the fracture tip whereby 𝑞̂ 𝐹𝑀 starts to linearly
decrease to zero. By virtue of 𝜕 𝑞̂ 𝐹𝑀 ∕𝜕 𝑥̂ = 0 for 𝑥̂ < 𝑥̂ 𝑞 , part of the first term on the right hand side of Equation 13 vanishes,
giving

𝑥̂
0 𝑥̂ 0 𝜕𝑝𝑓
1
𝐽𝑓 = 𝑤𝑝𝑓 𝑑𝑥̂ − 𝑞̂ 𝐹𝑀 𝑤 ̂
𝑑𝑥, (15)
𝑥̂ 0 − 𝑥̂ 𝑞 ∫𝑥̂ 𝑞 ∫0 𝜕 𝑥̂

where the fact that 𝑞𝐹𝑀 = 𝑞̂ 𝐹𝑀 on the fracture surfaces has been used. As a result, the use of the plateau 𝑞𝐹𝑀 function
ensures the first term of Equation 15 no longer evaluates fluid pressure between 0 < 𝑥̂ < 𝑥̂ 𝑞 .
PEZZULLI et al. 2171

The final modification involves substituting Equation 9 within Equation 15 to give

𝑥̂ 𝑥̂ 0
1 0
𝑞̂ 𝐹𝑀
𝐽𝑓𝐻𝐹𝑀 = 𝑤𝑝𝑓 𝑑𝑥̂ − 𝜇′ 𝑉 ̂
𝑑𝑥, (16)
𝑥̂ 0 − 𝑥̂ 𝑞 ∫𝑥̂ 𝑞 ∫0 𝑤

where fluid pressure gradients have been eliminated in exchange for apertures. The use of the Reynolds equation estab-
lishes the integral within the domain of HFM, and introduces a velocity and viscosity dependence. Combining Equa-
tions 10, 16 with 𝐽𝑟 from Equation 8 gives the complete 𝐽𝐻𝐹𝑀 -integral
( ) 𝐹𝑀 𝑥̂ 0 𝑥̂ 0 𝐹𝑀
𝜕𝑢𝑗 𝜕𝑞 1 𝑞̂
𝐽𝐻𝐹𝑀 = 𝜎𝑖𝑗 − 𝑊𝛿1𝑖 𝑑𝐴 + 𝑤𝑝𝑓 𝑑𝑥̂ − 𝜇′ 𝑉 𝑑𝑥̂ . (17)
∫𝐴∗ 𝜕𝑥1 𝜕𝑥𝑖 𝑥̂ 0 − 𝑥̂ 𝑞 ∫𝑥̂ 𝑞 ∫0 𝑤

where the use of the the plateau 𝑞𝐹𝑀 function (and therefore 𝑞̂ 𝐹𝑀 ) as shown in Figure 3 is implied. More detail on con-
structing the plateau function using finite elements is given in Section 4.3.
The dependence of 𝐽𝐻𝐹𝑀 on the propagation velocity may appear as a limitation of the method, since 𝑉 is unknown,
and arguably the quantity which we are trying to estimate. However, one must bear in mind that the 𝐽 and 𝐽𝐻𝐹𝑀 -integrals
assess the stability of a specific fracture geometry, that is, the geometry used at the time of the solution of the governing
equations. Specifying the fracture geometry implicitly specifies a propagation velocity. A trivial example would be if the
fracture geometry remains unchanged between time steps; the implied propagation velocity would be zero at each fracture
tip. Such an implied propagation velocity is what is needed within the 𝐽𝐻𝐹𝑀 to asses the stability of the current fracture
geometry. If the fracture is deemed to propagate upon using the 𝐽𝐻𝐹𝑀 -integral, then estimating the actual propagation
velocity of the fracture is desirable, and is tackled with the 𝐽𝐻𝐹𝑀 -algorithm in Section 6.
In the toughness-dominated regime, since the fluid pressure at the fracture tip is close to constant, the numerical solu-
tion for fluid pressure tends to be as accurate as the numerical solution for the aperture. Therefore the 𝐽 and 𝐽𝐻𝐹𝑀
integrals give identical estimates for the stress intensity factor 𝐾𝐼 . However, in the viscosity-dominated regime, large
fluid pressure drops at the fracture tip occur that can cause modest numerical errors in the fluid pressure. It will be
shown in Section 5 that such errors compromise the accuracy of the 𝐽-integral, but are avoided by design of the 𝐽𝐻𝐹𝑀
integral.

3.1 Asymptote-informed 𝑱𝑯𝑭𝑴 -integral

Evaluating the 𝐽𝐻𝐹𝑀 -integral in Equation 17 could still prove numerically difficult, the reason being that the third term
on the right hand side of Equation 17 involves the integration of a singular quantity 𝑤−1 . Integrating such a singular term
with methods such as Gauss–Legendre quadrature is susceptible to error given the relatively sparse number of integra-
tion points available to a quadratic interpolation. To avoid recourse to higher order methods or finer meshes, asymptotic
solutions for the fracture aperture can be used to evaluate the integrand containing 𝑤−1 analytically. The result detailed
𝐴
below is an asymptote-informed 𝐽𝐻𝐹𝑀 -integral (a 𝐽𝐻𝐹𝑀 -integral) that is valid throughout the toughness-viscous regime
of propagation. If only viscosity-dominated propagation is of concern, then the asymptote-informed 𝐽𝑉 -integral takes a
simpler form which is detailed in A.2.
An ideal candidate for the aperture at the fracture tip is the universal asymptote discovered by Dontsov and
Peirce23

( √ )1∕3
𝐾′ √ 3 𝑥̂
𝑤 = ′ 𝑥̂ 1 + 𝛽 , (18)
𝐸 𝓁𝑘𝑚

where the constant 𝛽 = 3.14735, and the term with the transition lengthscale 𝓁𝑘𝑚 incorporates the effects of viscous-
dominated propagation. The contribution from leak-off can also be included into Equation 18, but is omitted
here.24
2172 PEZZULLI et al.

Substituting Equation 18 and (14) into Equation 17 and carrying out the integration analytically gives the following
𝐴
𝐽𝐻𝐹𝑀 -integral optimised for propagation within the viscous-toughness regimes

( ) 𝐹𝑀 𝑥̂ 0
𝐴
𝜕𝑢𝑗 𝜕𝑞 1
𝐽𝐻𝐹𝑀 = 𝜎𝑖𝑗 − 𝑊𝛿1𝑖 𝑑𝐴 + 𝑤𝑝𝑓 𝑑𝑥̂
∫𝐴∗ 𝜕𝑥1 𝜕𝑥𝑖 𝑥̂ 0 − 𝑥̂ 𝑞 ∫𝑥̂ 𝑞
( )
3𝜇′ 𝑉𝐸 ′ √ 𝑟 [ ]
− 𝓁𝑘𝑚 𝑟𝑔(𝑥̂ 0 ) + (1 − 𝑟)𝑔(𝑥̂ 𝑞 ) − 1 + 𝑓(𝑥̂ 0 ) − 𝑓(𝑥̂ 𝑞 ) , (19a)
𝛽3𝐾′ 20𝛽 6 𝑥̂ 0

with the functions 𝑔, 𝑓 defined as

( √ )2∕3
𝑥̂ ( √ )
̂ =
𝑔(𝑥) 1 + 𝛽3 , ̂ 6𝛽 3 𝑥𝓁
̂ = 𝑔(𝑥)
𝑓(𝑥) ̂ 𝑘𝑚 − 5𝛽 6 𝑥̂ − 9𝓁𝑘𝑚 , (19b)
𝓁𝑘𝑚

where 𝑟 = 𝑥̂ 0 ∕(𝑥̂ 0 − 𝑥̂ 𝑞 ). Only the last term in Equation 17 has been modified by virtue of the analytical integration,
𝐴
which is expected to increase the accuracy of the integral. Fundamentally, the 𝐽𝐻𝐹𝑀 -integral analytically evaluates the
drop in the mechanical energy flux entering the fracture tip captured by the gradient in fluid pressure. Similarly to
the 𝐽𝐻𝐹𝑀 -integral in Equation 17, Section 19 avoids working with the fluid pressures in the vicinity encompassing the
𝐴
fracture tip (𝑥̂ < 𝑥̂ 𝑞 ). The performance of such a 𝐽𝐻𝐹𝑀 -integral in estimating the stress intensity factor is addressed in
Section 5.

4 FINITE ELEMENT IMPLEMENTATION DETAILS


𝐴
This section details the finite element model with which the numerical accuracy of the 𝐽, 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 integrals can
be evaluated. The aim is to achieve an accurate extraction of stress intensity factors 𝐾𝐼 without resorting to more compu-
tationally expensive or cumbersome procedures, hence a standard quadratic finite element implementation is carried out
and described below.

4.1 Numerical method

A discretisation procedure identical to Salimzadeh et al.57 has been developed for an impermeable rock and utilised herein.
Namely, Galerkin finite elements are employed for the spatial discretisation of the governing equations (Section 1), while
implicit Backward–Euler finite differences are used for the discretisation in time. A standard finite element formula-
tion with quadratic, isoparametric triangular elements is used.65 No contact algorithm is employed since the fracture is
expected to be open throughout the entirety of the simulation.
A coupled and non-linear system of equations is solved for the fluid pressure 𝑝𝑓 and rock displacements 𝒖 using a Picard
iterative procedure. The numerical model was developed within the Complex Systems Modelling Platform (CSMP++)70
where the functionality of representing fractures as explicit discontinuities within the finite element mesh was developed.
The linearised system is solved using the Algebraic Multigrid Methods for Systems (SAMG) Solver.71 Unless stated oth-
erwise, a minimum aperture 𝑤𝑚𝑖𝑛 = 1.5 × 10−4 m is imposed in order to guarantee a minimum hydraulic conductivity
of the fracture and thus the convergence of the solver. The smallest minimum width 𝑤𝑚𝑖𝑛 which ensures a robust con-
vergence of the SAMG solver is imposed, which is limited by the solver’s susceptibility to round-off errors. Such errors
depend on the coefficients of the lubrication equation and the floating-point representation, where a double-precision
format was used herein. The imposition of a minimum width will be shown to contribute to the overestimation of sin-
gular fluid pressures at the fracture tip (see Section 5). Consequently, such a minimum width is partially responsible
for the loss in accuracy of the original 𝐽-integral, but not the modified 𝐽𝐻𝐹𝑀 -integrals proposed herein (see Figure 5 in
Section 5).
PEZZULLI et al. 2173

F I G U R E 4 Finite element mesh with applied far field compressive stress 𝜎0 . By symmetry, half the volume flux 𝑄0 = 𝑄∕2 is injected.
The element size used along the crack face is ℎ = 0.125 m, with the location of the fracture when it is 10 m highlighted in red at 𝑥 = 0. The
size of the fracture at the start of the simulation is 50 cm

(A) (B)

F I G U R E 5 Analysis resulting from the numerical solution of the hydro-elastic equations combined with the analytically-informed
propagation in the viscosity-dominated regime ( = 0.36). (A) A snapshot in time 𝑡 = 18 s of the variation in the net pressure along the tip of
a 6 m fracture with ℎ = 0.125 m resolution for runs at different minimum conductivity 𝑤𝑚𝑖𝑛 . For the simulation run with 𝑤𝑚𝑖𝑛 = 1𝑒−4 meters,
𝐴
the normalised stress intensity factor 𝐾𝐼 ∕𝐾𝐴 is computed at every time step by the 𝐽, 𝐽𝐻𝐹𝑀 , and 𝐽𝐻𝐹𝑀 -integral and plotted in (B), along with the
𝐴
performance of the 𝐽𝐻𝐹𝑀 -integral when the remaining fluid pressures are evaluated using the analytical solution of Garagash and Detournay15

4.2 Problem geometry and finite element mesh

The classical KGD problem72,73 of a fracture propagating in plane-strain fed by a constant flux 𝑄 in an infinite medium
is solved with the finite element mesh shown in Figure 4. The symmetry of the problem allows for only half of the frac-
ture to be simulated, consequently injecting half of the flux 𝑄0 = 𝑄∕2. The injection point 𝑠 = 0 lies at the intersection
between the fracture and the left boundary of the model. Horizontal displacements along the whole of the left boundary
are fixed, while both horizontal and vertical displacements are fixed at a single point on the right boundary, as shown in
Figure 4. To allow the comparison with analytical solutions, the large domain size shown in Figure 4 was chosen, therefore
approximating an infinite body.74,75
2174 PEZZULLI et al.

Simulations were carried out at various mesh sizes. Figure 4 shows the mesh with the finest resolution along the fracture
ℎ = 0.125 m, which was used to generate the results in Section 5 and Section 7. As the fracture is anticipated to propagate
under pure Mode I loading, a pre-defined path ahead of the fracture is also refined, ensuring the same resolution ℎ along
the entirety of the fracture throughout the simulation. The fracture has a pre-existing size of 0.5 m and is simulated to
propagate a total of 𝐿 = 15 − 20 m, giving a final refinement ratio ℎ∕𝐿 ≲ 1∕120; refinement ratios smaller than 1∕20 are
typically expected to yield accurate results.
The values for the Young’s modulus 𝐸 = 17 GPa, Poisson’s ratio 𝜈 = 0.2, volume flux 𝑄 = 2𝑄0 = 8.52 × 10−4 m3 /s and
confining compressive stress 𝜎0 = 5 MPa were chosen and kept fixed for all simulations. The confining stress 𝜎0 is applied
on the horizontal boundaries and ensures the fluid lag within the fracture is negligible.63,64 The critical stress intensity
factor 𝐾𝐼𝑐 and the viscosity 𝜇 were selected according to the regime of propagation being studied.

4.3 Plateau 𝒒𝑭𝑴 function implementation

For all simulations, the domain of integration 𝐴∗ is chosen to contain two sets of element rings; an inner ring which
includes all the elements which share the node at the fracture tip, and an outer ring containing all elements that share a
node with the outer contour Γ0 . In the inner ring, 𝑞𝐹𝑀 = 1 is set on every node, while on the outer ring, 𝑞𝐹𝑀 is set to zero
on the nodes on the outer contour Γ0 , and mid-side nodes not lying on the outer contour Γ0 are set to 0.5. After-which,
linear finite elements are used to interpolate 𝑞𝐹𝑀 to the integration points for the integration.

5 NUMERICAL ANALYSIS OF THE 𝑱-INTEGRAL AND ITS HFM VARIATIONS

In order to asses the accuracy of the 𝐽-integral (Equation 8) and the variations introduced herein for HFM (Equations 17
and 19), a particular propagation algorithm is employed. The governing hydro-elastic equations are solved numerically
as described in Section 4, while the length of the fracture is prescribed to vary according to the analytical solution of a
fracture constantly propagating at 𝐾𝐼 = 𝐾𝐼𝑐 in the viscosity-dominated regime15 :

𝑄1∕2 𝐸 ′1∕6 2∕3


𝑙(𝑡) = (𝛾0 + ()𝛾1 ) 𝑡 , (20)
𝜇′1∕6

where 𝛾0 = 0.6152 and 𝛾1 = −0.17475, and the inclusion of 𝛾1 incorporates the effects
√ of a small finite toughness 𝐾𝐼𝑐 ≠ 0.
The small parameter () = 0.1076  3.16796 . Setting 𝜇 = 0.1Pas, 𝐾𝐼𝑐 = 1 MPa m gives a dimensionless toughness of
 = 0.36 suitable for viscosity-dominated propagation. Inverting Equation 20 gives the time step Δ𝑡 required to advance
the fracture tip by one element of length ℎ.
Prescribing the evolution in time of the fracture length as described by Equation 20 implies the stress intensity factor
𝐾𝐼 at each time step should be equal to the fracture toughness 𝐾𝐼𝑐 . In other words, the analytical solution 𝐾𝐴 for the stress
𝐴
intensity factor is 𝐾𝐴 = 𝐾𝐼𝑐 . At this point, the 𝐽, 𝐽𝐻𝐹𝑀 , 𝐽𝐻𝐹𝑀 -integrals are employed at each time step only to calculate
a numerical estimate for 𝐾𝐼 , thus not determining the propagation of the fracture. The propagation velocity 𝑉 used for
the HFM integrals is taken simply by dividing the distance propagated (in this case the element length ℎ) by the time
step Δ𝑡. The deviation of the ratio 𝐾𝐼 ∕𝐾𝐴 from unity details the normalised error incurred by the relevant integral in
estimating 𝐾𝐼 , and suffices to reveal the sources of error within the integrals. The error which the integrals incur if they
were to determine the length of the fracture are shown in Section 7 via adopting the propagation algorithms detailed in
Section 6.
Before the performance of the 𝐽 and 𝐽𝐻𝐹𝑀 integrals are evaluated, it is worth looking at the solution for the fluid pressure
in Figure 5A, which shows a snapshot of the fluid pressure profile at time 𝑡 = 18 s along the tip of a 6 m fracture. A
significant error is present within the numerical solution for the fluid pressure at the fracture tip. The analytical solution
for the fluid pressure in the viscosity-dominated regime entails a steep, non-linear pressure drop at the fracture tip; a
feature arising from the difficulty of driving viscous fluid through a vanishing aperture. Such singular nature is difficult
to capture numerically, and evidenced by the more gradual slope estimated by the numerical method (dashed lines). The
values set for the minimum aperture 𝑤𝑚𝑖𝑛 affect the extent of the pressure drop, where a smaller aperture allows for
a smaller hydraulic conductivity and therefore a greater pressure drop. However, significant errors remain in the fluid
pressure despite the use of minuscule minimum apertures reaching 𝑤𝑚𝑖𝑛 = 1𝑒−5 m. The numerical solution estimated
PEZZULLI et al. 2175

with 𝑤𝑚𝑖𝑛 = 1𝑒−4 m are used for the evaluation of the 𝐽 and 𝐽𝐻𝐹𝑀 integrals in the adjacent plot (Figure 5B). An important
remark is that the numerical fluid pressures are least accurate in the shaded regions of Figure 5A, where fluid pressure
values are taken for use within the integrals. In particular, the shaded region that encompasses the fracture tip describes
the fluid pressure values used by the 𝐽-integral, while the shaded region which avoids the fracture tip describes the values
𝐴
used by the 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 -integrals. Consequently, an increase in accuracy is expected when using the HFM integrals
to estimate 𝐾𝐼 .
The results for the evaluation of the 𝐽 and 𝐽𝐻𝐹𝑀 integrals can be seen in Figure 5B, which shows the normalised
stress intensity factor 𝐾𝐼 ∕𝐾𝐴 estimated by the integrals at every time step as the fracture propagated to a length of 20 m.
Figure 5B evidences that the errors in the fluid pressure 𝑝𝑓 are responsible for the majority of the overestimation of the
stress intensity factor 𝐾𝐼 . This can be deduced by observing that the 𝐽-integrals prediction for the stress intensity 𝐾𝐼 overes-
timates the analytical solution 𝐾𝐴 by approximately a factor of 3. Instead, exchanging fluid pressure for apertures by using
the 𝐽𝐻𝐹𝑀 reformulation reduces the error to a 2-fold overestimation of 𝐾𝐼 . The error in 𝐾𝐼 drops further if the asymptotic
𝐴
solution for the fracture aperture is used to integrate part of the integral analytical, as done by the 𝐽𝐻𝐹𝑀 -integral. Since the
errors in the numerical aperture at the fracture tip are small, such a result suggests the Gauss–Legendre integration pro-
𝐴
vides the main source of error when switching from the 𝐽𝐻𝐹𝑀 -integral to the 𝐽𝐻𝐹𝑀 -integral. Finally, a significant part of
𝐴
the remaining error can be eliminated if the remaining numerical fluid pressure values used by the 𝐽𝐻𝐹𝑀 -integral (second
term on the right hand side of Equation 19a) are exchanged with the three-term analytical solution for the fluid pressure
provided by Garagash and Detournay.15 The use of the Garagash and Detournay analytical solution is not relevant for prac-
tical numerical applications, but it serves the purpose of showing that the 𝐽-integral is theoretically capable of estimating
stress intensity factors in the viscous-dominated regime, but its capability is hindered by the accuracy of the numerical
solution for fluid pressure.
The error observed in the 𝐽-integral does not depend on whether the 𝐽-integral is applied where the toughness-
asymptote is valid. This can be seen by noting the transition lengthscale varied from 0.01 < 𝓁𝑘𝑚 < 0.35 throughout the
simulation. Hence, the resolution used of ℎ = 0.125 meters does not capture the toughness asymptote, which exists at
ratios smaller than ℎ∕𝓁𝑘𝑚 ≲ 1𝑒−3 .17 Consequently, the elimination of the majority of the error from the 𝐽-integral evi-
dences the fact that the performance of the 𝐽-integral is not dependent on the resolution of the toughness asymptote.
Such a fact is theoretically established by the derivation of the 𝐽-integral from purely momentum balance considerations.45
Hence, the 𝐽-integral can theoretically compute the stress intensity factor in the toughness-viscous regime, so long as the
numerical solution for the stress, displacement and fluid pressure are accurate at the fracture tip.
While the 𝐽-integral can be validly applied throughout the toughness-viscous propagation regime, its raw form Equa-
tion 8 is not advised for HFM where singular fluid pressures exist. The spatial steps ℎ = 0.125m and time steps Δ𝑡 =
0.1 − 0.5s used in this section are already restrictive for practical applications on the reservoir scale. With higher spatial–
temporal resolution,60 or perhaps alternative numerical methods like the finite volume method, the numerical fluid pres-
sure at the fracture tip may be better approximated, thus eliminating the source of error. However, such approaches could
still prove prohibitive for large-scale applications. A commonly-used alternative to handling the singular tip pressures
involves prescribing the numerical solution at the fracture tip to behave according to the tip aperture asymptotics.32,34,35,76
Consequently, the singular fluid pressures are adequately captured, but come at the cost of requiring bespoke coupling
𝐴
schemes.35,76 Instead, by applying the reformulated integrals 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 (see sections 6, 7), the propagation of
a hydraulic fracture can be accurately simulated despite the inaccurate fluid pressures arising in the numerical solu-
tion. A similar philosophy has been used with the 𝐽-integral to handle inaccurate contact tractions over the fracture
surfaces.77

6 AN IMPLICIT FRONT-TRACKING 𝑱𝑯𝑭𝑴 -ALGORITHM

While estimating the energy release rate informs us on whether the fracture is propagating, it does not provide any infor-
mation on the possible location of the new fracture front. To converge on the location for the fracture front during a
time step, an implicit front-tracking algorithm is required. The following 𝐽𝐻𝐹𝑀 -algorithm achieves this goal by iteratively
repeating the following three steps until convergence on the fracture-front is achieved:

𝐴
1. Calculate the energy release rate at a current trial fracture front (the 𝐽𝐻𝐹𝑀 -integral is used here)
2. Invert the energy release rate to predict the propagation velocity of the fracture
2176 PEZZULLI et al.

𝐴
F I G U R E 6 Implicit front-tracking 𝐽𝐻𝐹𝑀 -algorithm that uses the 𝐽𝐻𝐹𝑀 -integral to calculate the stress intensity factor and invert it to
iterate on the fracture boundary. The procedure to find the solution for the fracture length 𝓁𝑛 , fluid pressure 𝐩𝑛𝑓 and displacement of the rock
𝐮𝑛 is shown for the time step 𝑛

3. Modify the time step such that the fracture is expected to propagate one element at a time

𝐴
The previous section detailed how step 1 can be performed using the 𝐽𝐻𝐹𝑀 -integral throughout the toughness-viscous
regime of propagation. This section describes how such a process can be integrated with steps 2 and 3 to construct a
𝐽𝐻𝐹𝑀 -algorithm that implicitly converges on the fracture front. It is worth clearly stating that modification of the global
time step is not a procedure that can be generalised to handle multiple fracture tips, or 3-dimensions, since different
tip propagation velocities and mesh sizes will require different modifications of the time step. Instead, for more general
settings, modification of the underlying mesh, or prescribing the tip asymptotics,19,24,31 is needed to ensure convergence.

6.1 Numerical Algorithm

An overview of the implicit front tracking scheme used herein to iterate on the fracture length is shown in Figure 6
for a time step 𝑛. Two iterative loops are required at each time step. An inner loop implicitly solves via Picard iteration
the coupled non-linear hydro-elastic equations for the displacement of the solid 𝐮𝑛,𝑖 and the fluid pressure 𝐩𝑛,𝑖 𝑓
for a
particular iteration 𝑖 at a current trial fracture length 𝓁𝑛,𝑖 . The procedure is repeated for each iteration 𝑖 of the outer loop
until convergence on the final fracture length 𝓁𝑛 is reached. Since the hydro-elastic equations are solved assuming a trial
PEZZULLI et al. 2177

F I G U R E 7 The various fracture fronts


relevant for the iteration 𝑖 and time step 𝑛 of the
implicit front-tracking 𝐽𝐻𝐹𝑀 -algorithm, and their
connection to the coordinates used by the
𝐴 𝐴
𝐽𝐻𝐹𝑀 -integral. The estimate for 𝐽𝐻𝐹𝑀 at a trial
fracture front 𝓁𝑛,𝑖 can be used to solve for 𝑥∗ . The
old fracture front 𝓁𝑛−1 need not coincide with 𝑥̂ 𝑞 ,
but is the case when propagation occurs in one
element increments, and 2 element rings are used
𝐴
with the 𝐽𝐻𝐹𝑀 -integral

position for the fracture length 𝓁𝑛,𝑖 , a trial propagation velocity 𝑉 𝑛,𝑖 = (𝓁𝑛,𝑖 − 𝓁𝑛−1 )∕Δ𝑡𝑛,𝑖 can always be formulated since
the fracture length at the previous time step 𝓁𝑛−1 is also known, along with the time step Δ𝑡𝑛,𝑖 used for the current iteration.
For the first iteration 𝑖 = 0, the fracture is assumed to be at rest and therefore the fracture length at the previous time
𝐴
step is chosen as the trial fracture length 𝓁𝑛,0 = 𝓁𝑛−1 . At this point, the 𝐽𝐻𝐹𝑀 -integral is applied as written in Section 19 to
𝑛,𝑖
calculate the stress intensity factor 𝐾𝐼 with propagation velocity 𝑉 = 0 for a fracture at rest. Subsequently, if the fracture
is not deemed to propagate (𝐾𝐼 < 𝐾𝐼𝑐 ), no subsequent iteration on the fracture length is needed, and the algorithm can
move on to the next time step. Alternatively, if propagation is predicted (𝐾𝐼 > 𝐾𝐼𝑐 ), iteration on the fracture length is then
enacted (𝑖 = 1) and the length of the fracture is incremented by one element 𝓁𝑛,1 = 𝓁𝑛,0 + ℎ. At this point, the fracture is
considered to be propagating (𝓁𝑛,𝑖 > 𝓁𝑛−1 ).
𝐴
For the propagating fracture, 𝐽𝐻𝐹𝑀 is calculated using Section 19 using the new trial propagation velocity 𝑉 𝑛,𝑖 corre-
𝐴
sponding to the new trial configuration 𝓁𝑛,𝑖 . Once 𝐽𝐻𝐹𝑀 is calculated, a location for the new position of the fracture-front
𝓁 can be estimated by inverting the implicit equation

𝐴
𝐽𝐻𝐹𝑀 = 𝑓(𝑥 ∗ , 𝑉(𝑥∗ ); 𝐾𝐼𝑐 ) (21)

for 𝑥∗ , which gives the distance to the unknown fracture length 𝓁 = 𝓁𝑛,𝑖 + 𝑥 ∗ − 𝑥̂ 𝑞 . A schematic of the relevant coordinate
and fracture lengths for the inversion is shown in Figure 7. The unknown 𝑥∗ represents the length 𝑥̂ 𝑞 would have to
be for the fracture to propagate at the fracture toughness 𝐾𝐼𝑐 . This also implies that the right hand side of Equation 21
depends on the fracture toughness 𝐾𝐼𝑐 and not the estimated stress intensity factor 𝐾𝐼 . The propagation velocity 𝑉(𝑥∗ )
is a function of the unknown 𝑥∗ which can be written as 𝑉 = (𝓁 − 𝓁𝑛−1 )∕Δ𝑡𝑛,𝑖 , where 𝓁 was shown above to depend on
𝑥∗ . In the particular case herein, the fracture propagates one element of length ℎ per time step, therefore 𝑥̂ 𝑞 coincides
𝐴
with the old fracture length 𝓁𝑛−1 since two element rings are used with the 𝐽𝐻𝐹𝑀 -integral. Such a configuration gives
𝓁=𝓁 𝑛−1 + 𝑥 and 𝑉 = 𝑥 ∕Δ𝑡 , as can be seen in Figure 7. The implicit equation Equation 21 is solved for 𝑥 ∗ using
∗ ∗ 𝑛,𝑖
𝐴
the bisection method. The inversion of 𝐽𝐻𝐹𝑀 to obtain the propagation velocity 𝑉 is a novel alternative to the aperture-
based approaches in estimating the velocity of propagation in HFM.20,24,34,78 The details of the inversion are described in
Section 6.3.
Once an estimate for the propagation velocity 𝑉 is obtained, it is compared to the the trial velocity the fracture has
travelled in the simulation 𝑉 𝑛,𝑖 . Convergence of the fracture algorithm is accepted if

|𝑉 − 𝑉 𝑛,𝑖 |
| | < 𝜀𝑣 , (22)
𝑉 𝑛,𝑖

whereby 𝜀𝑣 is a prescribed tolerance set to 𝜀𝑣 = 0.005 for all simulations. If convergence is not achieved, the time step is
modified to scale the estimated fracture front 𝓁 to match with the trial fracture length 𝓁𝑛,𝑖 . By using the estimated propa-
gation velocity 𝑉, the time step Δ𝑡∗ = ℎ∕𝑉 can be estimated. The final time step used for the next iteration is then obtained
via taking the weighted average Δ𝑡𝑛,𝑖+1 = 𝛼Δ𝑡∗ + (1 − 𝛼)Δ𝑡𝑛,𝑖 . The weighted average is useful to ensure a monotonic con-
vergence to the correct fracture front, however setting 𝛼 = 1.0 gives the fastest convergence rate and is used herein. Instead
of modifying the time step, re-meshing of the fracture tip could be employed by modifying the propagation increment ℎ.
2178 PEZZULLI et al.

6.2 Tracking the fracture front using Paris-type convergence iteration

For comparison with the implicit front-tracking 𝐽𝐻𝐹𝑀 -algorithm, a front-tracking scheme inspired by the Paris-type con-
vergence procedure is implemented for the classical 𝐽-integral.79 In this case,

( )𝑏
𝐾𝐼𝑐
Δ𝑡∗ = Δ𝑡𝑛,𝑖 , (23)
𝐾𝐼

and a weighted average between Δ𝑡∗ and Δ𝑡𝑛,𝑖 is taken to get the time step for the next iteration Δ𝑡𝑛,𝑖+1 . The weighted
average parameter chosen was 𝛼 = 0.9, while the exponent 𝑏 was 0.3 and 0.03 in toughness and viscosity dominated
propagation. Such a convergence procedure is ad-hoc, where more advanced ‘static’ approaches entail using linear
interpolation.60,80 Nonetheless, the approach implemented herein is sufficient as it serves the purpose of achieving a
converged result for comparison with the 𝐽𝐻𝐹𝑀 -algorithm.

6.3 Extracting the propagation velocity from the energy release rate
𝐴
Once 𝐽𝐻𝐹𝑀 has been evaluated, the result can be inverted to estimate a distance to the fracture front, and by extension, the
propagation velocity of the fracture. The process involves inverting Equation 21 for the unknown distance to the fracture
𝐴
tip 𝑥∗ . Choosing 𝑥 ∗ as the distance from the tip to the start of the last ring of elements 𝑥̂ 𝑞 allows the 𝐽𝐻𝐹𝑀 -integral from
Equation 17 to be formulated as

𝑥∗ +𝑎 𝑥∗ +𝑎
𝐴 1 𝑞̂ 𝐻𝐹𝑀
𝐽𝐻𝐹𝑀 = 𝐽𝑟 (𝑥 ∗ ) + 𝑤𝑝𝑓 𝑑𝑥̂ − 𝜇′ 𝑉 𝑑𝑥̂ . (24)
𝑎 ∫𝑥 ∗ ∫0 𝑤

𝐴 𝐴
where 𝐽𝐻𝐹𝑀 is the numerical evaluation of the energy release rate done using the 𝐽𝐻𝐹𝑀 -integral for the current fracture
𝑛,𝑖
length 𝓁 . The constant 𝑎 = 𝑥̂ 0 − 𝑥̂ 𝑞 can be understood by recalling the geometry of the contours relevant to the 𝐽-
integral shown in Figure 3. The challenge lies in revealing the dependence on 𝑥∗ of the aperture 𝑤, fluid pressure 𝑝𝑓 , and
mechanical energy flux entering the fracture tip 𝐽𝑟 (defined in Equation 8).
A combination of numerical and asymptotic solutions are used to specify such a dependence. For the aperture 𝑤, the
asymptote shown in Equation 18 discovered by Dontsov and Peirce23 is used. Instead, for the mechanical energy flux term
an empirical relationship was found that follows the dependence

𝑥̂ 𝑞
𝐽𝑟 = 𝐽𝑟𝑖 , (25)
𝑥∗

where 𝐽𝑟𝑖 is the numerically computed 𝐽𝑟 obtained at the current trial fracture length 𝓁𝑛,𝑖 . The functional relationship
assumed in Equation 25 was calibrated via fitting the numerically observed drop in 𝐽𝑟 for increases in the trial fracture
length 𝓁𝑛,𝑖 . Introducing Equation 25 solely serves the purpose of ensuring convergence of the fracture-front iteration,
and does not change the final result. This is recognised by noting that convergence on the final fracture length implies
𝑥 ∗ → 𝑥̂ 𝑞 , giving 𝐽𝑟 → 𝐽𝑟𝑖 as expected.
For the asymptotic fluid pressure, the following novel semi-analytical solution

[( )1∕3 ( √ )]𝑥̂ 0
3𝜇′ 𝑉𝐸 ′2 𝓁𝑚𝑘 2 2 5 1 𝓁𝑚𝑘
̂ =
𝑝𝑓 (𝑥) 𝜆𝑝𝑓𝑖 (𝑥̂ 0 ) + 2 𝐹1 , ; ;− , (26)
𝛽 2 𝐾 ′2 𝑥 3 3 3 𝛽3 𝑥
𝑥̂

was found, and is valid for a plane-strain fracture propagating throughout the entire toughness-viscous regime of
propagation. The function 2 𝐹1 (⋅ , ⋅ ; ⋅ ; ⋅) represents the gaussian hypergeometric function. The quantity 𝑝𝑓𝑖 (𝑥̂ 0 ) is the
numerical fluid pressure evaluated at 𝑥̂ 0 , and denoted simply 𝑝𝑓𝑖 henceforth. The analytical bracketed term in Equa-
̂ Similar to Equa-
tion 26 is the result of a definite integral, and therefore the variable 𝑥 must be evaluated at 𝑥̂ 0 and 𝑥.
∗ 1∕3
tion 25, a convergence parameter 𝜆 = (𝑥̂ 𝑞 ∕𝑥 ) has been added to ensure the convergence of the scheme. Its role is to
PEZZULLI et al. 2179

represent the drop in the absolute value of the fluid pressure due to a change in length of the fracture. As the algorithm
converges to the correct time-step, 𝑥∗ → 𝑥̂ 𝑞 and 𝜆 converges to 1, which reflects the fact that the analytical and numer-
ical estimates for fluid pressure at 𝑥̂ 0 must match as convergence is reached. The derivation of Equation 26 is detailed
in B.
Substituting Equation 25, Equation 26, and Equation 18 into Equation 24 gives the final implicit equation

𝑥̂ 𝑞
𝐴
𝐽𝐻𝐹𝑀 = 𝐽𝑟𝑖 + 𝐽𝑤𝑝 (𝑥 ∗ ) − 𝐽𝜇 (𝑥 ∗ ), (27a)
𝑥∗

where
( )
3𝜇′ 𝑉𝐸 ′ √ 1
𝐽𝜇 (𝑥 ∗ ) = 𝓁𝑘𝑚 ̂
𝑟𝑔(𝑥 ∗ + 𝑎) + (1 − 𝑟)𝑔(𝑥
̂ ∗) − 1 +
[𝑓(𝑥 ∗ + 𝑎) − 𝑓(𝑥 ∗ )] , (27b)
𝛽3𝐾′ 20𝑎𝛽 6

( )1∕3 √
⎡ ⎡ ⎛ ⎞⎤
1 𝓁 𝑘𝑚 2 2 5 1 𝓁𝑘𝑚 ⎟⎥
𝐽𝑤𝑝 (𝑥 ∗ ) = ⎢(𝑊(𝑥 ∗ ) − 𝑊(𝑥 ∗ + 𝑎))⎢𝜆 𝑝𝑓𝑖 + 𝑀 2 𝐹1
⎜ , ; ;−
𝑎⎢ ⎢ 𝑥̂ 0 ⎜ 3 3 3 𝛽3 𝑥̂ 0 ⎟⎥
⎣ ⎣ ⎝ ⎠⎦
( √ ) ]
𝑥∗ +𝑎 ( )1∕3 √
𝐾′ 𝓁𝑘𝑚 2 2 5 1 𝓁𝑘𝑚
−𝑀 ′ 𝑥̂ 𝑔(𝑥)
̂ 2 𝐹1 , ; ;− 𝑑𝑥̂ , (27c)
𝐸 ∫𝑥∗ 𝑥̂ 3 3 3 𝛽3 𝑥̂

where the functions 𝑔, 𝑓 are defined as in Equation 19b, while the lumped parameter 𝑀 and function 𝑊 are defined as

3𝜇′ 𝑉𝐸 ′2 3𝐾 ′ 𝓁 ( √ )
𝑘𝑚
𝑀= , 𝑊(𝑥 ∗ ) = 𝑔(𝑥 ∗ )2 12𝛽 3 𝑥 ∗ 𝓁
𝑘𝑚 − 14𝛽 6 𝑥 ∗ − 9𝓁
𝑘𝑚 . (27d)
𝛽 2 𝐾 ′2 70𝛽 9 𝐸 ′

and 𝑟̂ = (𝑥∗ + 𝑎)∕𝑎 and 𝛽 = 3.14735 as before. It should be noted that 𝓁𝑘𝑚 is dependent on 𝑥∗ via its dependence on
the unknown propagation velocity 𝑉. The algebra is admittedly cumbersome, and an unavoidable result of attempting to
express the 𝐽𝐻𝐹𝑀 -integral in an analytical form. The use of the numerical estimates 𝐽𝑟𝑖 and 𝑝𝑓𝑖 within Section 27 is required
due to the absence of asymptotic solutions valid throughout the toughness-viscous regime of propagation. The bisection
method is used to solve Section 27 for the distance to the tip 𝑥∗ , whereby the integral term in Equation 27c is evaluated
numerically using Simpson’s integration with 𝑁 = 1000 intervals. The gaussian hypergeometric function 2 𝐹1 (⋅ , ⋅ ; ⋅ ; 𝑧) is
evaluated using the GSL library,81 where a Pfaff transformation is used to evaluate the integral for 𝑧 < −1. Occasionally,
𝐴
over-propagation of the fracture leads to negative values predicted for 𝐽𝐻𝐹𝑀 . In this case, the left hand side of Equation 27a

is set to zero before solving for 𝑥 , which represents an upper bound of the distance travelled by the fracture tip.
The advantage obtained in using Section 27 is that the propagation velocity can be extracted throughout the entire
toughness-viscous regime of propagation from numerical estimates of the energy release rate 𝐺 (which can be obtained
𝐴
using the 𝐽𝐻𝐹𝑀 -integral). Such a procedure is a novel energy-based alternative to methods which rely solely on inverting
the fracture aperture 𝑤 to extract the propagation velocity.19,20,24,78 Furthermore, the inversion of the energy release rate
is less susceptible to local errors in the aperture 𝑤, a common feature in the toughness-dominated regime. The relevant
analogy in fracture mechanics would be the comparison between the 𝐽-integral and the displacement correlation method.

7 RESULTS AND DISCUSSION

In this section, we evaluate the performance of the implicit front-tracking 𝐽𝐻𝐹𝑀 -algorithm described in Section 6 in com-
bination with the finite element model described in Section 4. The accuracy of the method is tested by simulating the
propagation of a KGD fracture at the extreme ends of the toughness and viscosity-dominated regimes. Furthermore, its
performance is investigated for various discretisations of the fracture. Consequently, the accuracy of the method at coarser
resolutions is assessed; a feature of great relevance when running reservoir scale simulations. In all results, percent errors
of a quantity 𝜙, with analytical solution 𝜙𝑎 , are defined as 𝜀̃𝜙 = 100(𝜙 − 𝜙𝑎 )∕𝜙𝑎 , where 𝜀𝜙 = |̃𝜀𝜙 | is the absolute value of
the percent error 𝜀̃𝜙 .
2180 PEZZULLI et al.

F I G U R E 8 Predictions and errors in the viscosity-dominated regime for the front-tracking 𝐽𝐻𝐹𝑀 algorithm compared with the classical
𝐽-integral that uses the Paris-type iteration to converge on the fracture front

7.1 Viscosity-dominated propagation

The analytical solutions for the fracture half-length 𝓁, crack mouth aperture 𝑤𝑐 = 𝑤(𝑠 = 0, 𝑡), and crack mouth net pres-
sure 𝑝 = 𝑝𝑓 (𝑠 = 0, 𝑡) − 𝜎0 for viscosity-dominated propagation are15

( )1 ( )1∕6 ( )1∕3
𝐸 ′ 𝑄3 6 2 𝜇 ′ 𝑄3 𝐸 ′2 𝜇′
𝓁(𝑡) = 𝛾0 𝑡3 , 𝑤𝑐 (𝑡) = 𝛾0 Ω0 𝑡1∕3 , 𝑝(𝑡) = Π0 , (28)
𝜇′ 𝐸′ 𝑡

where the dimensionless quantities 𝛾0 = 0.6152, Ω0 = 1.84475 √ and Π0 = 0.5469. A dimensionless toughness  = 0.036 is
simulated by setting the fracture toughness 𝐾𝐼𝑐 = 0.1 MPa m, and the viscosity to 𝜇 = 0.1 Pas. In doing so, the transition
lengthscale stays below 𝓁𝑘𝑚 < 1 × 10−6 m throughout the simulation; a distance significantly smaller than the resolution
at the fracture tip ℎ = 0.125 m. The remaining physical parameters are as discussed in Section 4.
The results in the viscosity-dominated regime for the 𝐽𝐻𝐹𝑀 -algorithm are compared in Figure 8 to the 𝐽-integral that
uses the Paris-type algorithm (Section 6.2). The variation in time of the quantities in Equation 28 is shown on the left hand
side of Figure 8, while their associated percent error is displayed on the right hand side. The absolute value of the percent
error in the fracture half length, crack mouth aperture, and fluid pressure are denoted 𝜀𝑙 , 𝜀𝑤 , 𝜀𝑝 , respectively.
The results in Figure 8 show that the classical 𝐽-integral consistently over-estimates the fracture length in the viscosity-
dominated regime. This can be seen by the percent error in the fracture length remaining above 𝜀𝑙 > 7% for the majority
of the simulation. Such a result is a direct symptom of the over-estimation of the stress intensity factor arising from the
over-estimation of the fluid pressure at the fracture tip observed in the analysis of Section 5. The overestimation of the
fracture length subsequently causes errors in the fluid pressure and the aperture profile at the crack mouth averaging
PEZZULLI et al. 2181

F I G U R E 9 Prediction and error in the fracture length for the toughness-dominated regime for the front-tracking 𝐽𝐻𝐹𝑀 -algorithm and
𝐴
the Paris-type iteration converging on the 𝐾𝐼 = 𝐾𝐼𝑐 using the 𝐽 or 𝐽𝐻𝐹𝑀 -integral

around 5%. Hence, the use of the 𝐽-integral in the viscosity-dominated regime transforms local errors at the fracture tip
to global errors present throughout the fracture.
On the other hand, the 𝐽𝐻𝐹𝑀 -algorithm displays an excellent accuracy in predicting propagation in the viscosity-
dominated regime. The success of the method is demonstrated in Figure 8 by the error in the fracture half-length remaining
𝐴
below 𝜀𝑙 < 1% for the majority of the simulation. Such a feature is a consequence of the increased accuracy of the 𝐽𝐻𝐹𝑀 -
integral demonstrated in Figure 5, which is used by the 𝐽𝐻𝐹𝑀 -algorithm to compute the stress intensity factor 𝐾𝐼 . Although
the error analysis in Figure 5 uses a different dimensionless toughness to the results in Figure 8, one may conclude that
the energy release rate provides meaningful information to accurately locate the fracture front in the viscosity-dominated
regime, despite errors in the stress intensity factor calculation reaching 10 − 100 percentage points (i.e., see Figure 5B for
𝐴
the 𝐽𝐻𝐹𝑀 -integral). Hence, larger errors in the calculation of the stress intensity factor can generally be accepted in the
viscous-dominated regime compared to the toughness-dominated regime. Yet, errors in the stress intensity factor com-
putation may be large enough as to compromise the accuracy of the location of the fracture front, as is the case for the
classical 𝐽−integral. Finally, Since ℎ ≫ 𝓁𝑚𝑘 , the performance of the 𝐽𝐻𝐹𝑀 −algorithm reinforces the idea that energy-
based methods have the capacity to achieve accurate results independently to the transition lengthscale.

7.2 Toughness-dominated propagation

The performance of the 𝐽𝐻𝐹𝑀 -algorithm must also be tested for the other extreme case of toughness-dominated prop-
agation. Only the results for the fracture half-length 𝓁 and its associated percent error 𝜀𝑙 are shown in Figure 9 for the
toughness-dominated regime, as they are most indicative of the performance of the method. The analytical solution for
the fracture half-length 𝓁 is 16

( )2∕3
𝐸′𝑄
𝓁(𝑡) = 𝛾0 𝑡2∕3 , (29a)
𝐾′

where 𝛾0 = 2𝜋−2∕3 . A dimensionless toughness  = 4.13 simulates the propagation of hydraulic √ fractures in the
toughness-dominated regime. This was simulated by setting 𝜇 = 1 × 10−4 Pas and 𝐾𝐼𝑐 = 2 MPa m.
The well-known accuracy of the classical 𝐽-integral is observed in the toughness-dominated regime. Such a result can
be seen in Figure 9 by the errors in the fracture length 𝜀𝑙 < 1% and comes as no surprise as the performance of 𝐽-integral
in such a regime has been established for decades.38 Hence, the result also serves as a verification of the 𝐽-integral’s
implementation, along with the finite element model described in Section 4.
The 𝐽𝐻𝐹𝑀 -algorithm slightly underestimates the fracture length in the toughness-dominated regime. More specifically,
an error averaging around 𝜀𝓁 ∼ 3% can be observed in the fracture half-length in Figure 9. Such an error is a result of the
choice of convergence parameter used by the 𝐽𝐻𝐹𝑀 -algorithm, whereby errors are introduced when locating the fracture
front using Section 27. In an attempt to prove the source of the error, the 𝐽𝐻𝐹𝑀 -integral is applied to predict 𝐾𝐼 , while the
Paris-type iteration in Section 6.2 is used to converge on 𝐾𝐼𝑐 . Such a method results in a prediction of the fracture length
that is virtually indistinguishable from the classical 𝐽-integral. Hence, essentially identical values for the stress intensity
2182 PEZZULLI et al.

F I G U R E 1 0 Snapshots in time of the aperture profile of a hydraulic fracture predicted by the 𝐽-integral and 𝐽𝐻𝐹𝑀 -algorithm in the
toughness and viscosity-dominated propagation regimes. The analytical solution is taken from the data repository provided in Dontsov82

𝐴
factor are predicted by the 𝐽 and 𝐽𝐻𝐹𝑀 -integrals in the toughness-dominated regime, and the source of the error in the
𝐽𝐻𝐹𝑀 −algorithm must lie when locating the fracture front. The authors do not recommend changing the convergence
criteria to the stress intensity factor, as that may incur similar errors in the viscosity-dominated regime. Such a feat is done
herein for the sole purpose of proving the source of the error. In practice, the error conferred by the 𝐽𝐻𝐹𝑀 −algorithm is
deemed acceptable, particularly since a fast and robust convergence on the fracture front is also guaranteed by the method,
as will be shown in Section 7.4.

7.3 Aperture Snapshots

A comparison of the aperture profiles predicted by the 𝐽−integral and 𝐽𝐻𝐹𝑀 −algorithm are shown in Figure 10 for the
toughness and viscosity-dominated propagation regimes. Such a figure summarises the accuracy previously discussed of
the two methods. Particularly, the over-prediction of the fracture length of the 𝐽−integral can be seen for all snapshots 𝑡 ∼
20, 42, 70 s in the viscosity-dominated regime. Consequently, smaller apertures result at the middle of the fracture 𝑥 = 0.
Alternatively, the 𝐽𝐻𝐹𝑀 −algorithm is shown to achieve an excellent match for the length and aperture of the fracture at
the various snapshots. On the other-hand, in the toughness-dominated regime, both methods achieve a good match with
the analytical solution of Dontsov82 at the times 𝑡 ∼ 30, 71, 120 s; however, the 𝐽𝐻𝐹𝑀 −algorithm slightly under-estimates
the fracture length compared to the 𝐽−integral.

7.4 Resolution requirements and convergence speed

The performance of the method at coarse resolutions is paramount to run reservoir scale simulations. To this end, the
accuracy for a certain element-size-to-fracture-length ratio ℎ∕𝓁 must be evaluated, along with the number of iterations
needed to converge on the fracture length. Hence, simulations similar to the preceding section were run, however the
percent error at each time step in the fracture length 𝜀̃𝑙 was now recast in terms of ℎ∕𝓁 and shown in Figure 11A and
Figure 11B. In addition, the number of iterations required to converge on the fracture front for the 𝐽𝐻𝐹𝑀 -algorithm, and
the Paris-type iteration are plotted in Figure 11C and Figure 11D. A fixed element length ℎ = 0.25 m along the fracture
PEZZULLI et al. 2183

(A) (B)

(C) (D)

F I G U R E 1 1 (A, B) Variation in the error in the fracture length 𝜀̃𝑙 for different element-length-to-fracture-length ratios ℎ∕𝓁. The element
length ℎ = 0.25 m was kept fixed as the fracture propagated throughout the simulation. (C, D) Number of iterations required on the fracture
front for the 𝐽𝐻𝐹𝑀 -algorithm, and the Paris-type iteration. Results are reported for the toughness-dominated ( = 4.13) and
viscosity-dominated ( = 0.036) regimes of propagation

was chosen, while the fracture length 𝓁 propagates from 0.5 to 15 m throughout the simulation. Consequently, as the
simulation progresses, some errors in the fracture length accumulate from the temporal discretisation, leading to results
at small ℎ∕𝓁 having marginally higher accumulated errors from the temporal discretisation. Nonetheless, such an effect
is small compared to the spatial discretisation of the fracture tip, since simulations run for various fixed element lengths
ℎ = (0.125, 0.5, 1, 2, 4, 8) m showed similar results, despite time steps ranging from Δ𝑡 ∼ (0.1 − 0.3) 𝑠 for ℎ = 0.125 𝑚, to
Δ𝑡 ∼ (10 − 80) 𝑠 for ℎ = 8 𝑚. The results in Figure 11 are classified in terms of the toughness ( = 4.13) and viscosity
( = 0.036) dominated regimes.
In the toughness dominated regime, Figure 11A shows that both the 𝐽𝐻𝐹𝑀 -algorithm and the 𝐽-integral using the Paris-
type iteration maintain their accuracy at coarser resolutions of the fracture length. With 5 elements discretizing the fracture
(ℎ∕𝓁 = 0.2), both methods report errors 𝜀𝑙 < 5%, with the 𝐽-integral algorithm showcasing the lowest error at around
𝜀̃𝑙 ∼ 3%. Using 10 elements (ℎ∕𝓁 = 0.2) or more along the fracture ensures the 𝐽-integral has 𝜀̃𝑙 ≲ 2% error, while the 𝐽𝐻𝐹𝑀 -
algorithm undershoots the fracture length by less than 4%. As mentioned in the preceding section, such a difference can
be eliminated by converging on the fracture toughness with the 𝐽𝐻𝐹𝑀 -integral, instead of the propagation velocity. Overall,
the accuracy of both methods are promising for coarse-scale simulations in the toughness-dominated regimes.
In the viscosity-dominated regime, Figure 11B shows that the performance of the classical 𝐽-integral quickly deteriorates
at coarser resolutions. Indeed, the error in the fracture length remains above 𝜀̃𝑙 > 10% for most refinement ratios ℎ∕𝓁 sim-
ulated. Significantly higher errors 𝜀̃𝑙 ≳ 30% are observed when fewer than 10 elements (ℎ∕𝓁 = 0.1) discretize the fracture
length. It is clear that the classical 𝐽-integral becomes quickly prohibitive for viscosity-dominated propagation at coarse
resolutions. In contrast, the 𝐽𝐻𝐹𝑀 -algorithm undershoots the error in the fracture length by approximately |̃𝜀𝑙 | ∼ 5% when
2184 PEZZULLI et al.

five elements resolve the fracture tip, and by |̃𝜀𝑙 | ≲ 2% when using more than 10 elements. Overall, the 𝐽𝐻𝐹𝑀 -algorithm
maintains its accuracy for coarser scale resolutions in the viscosity-dominated regime.
The number of iterations required to converge on the fracture front is an important metric which evaluates the computa-
tional cost of the algorithm. Figure 11C and Figure 11D show that the 𝐽𝐻𝐹𝑀 -algorithm converged on the fracture length with
an average of approximately 5 iterations in the toughness-dominated regime, and 4 iterations in the viscosity-dominated
regime. The number of iterations also decrease in time throughout the simulation for both regimes, since the previous
time step provides a better initial guess when the propagation velocity varies less between time steps. Overall, the 𝐽𝐻𝐹𝑀 -
algorithm presents a dramatic speed-up compared to the Paris-type inspired approach employed for the classical 𝐽-integral,
which averages around 15 and 11 iterations in the toughness and viscosity-dominated regimes, respectively. Indeed, the
convergence rate of the 𝐽𝐻𝐹𝑀 -algorithm is on-par with other state-of-the-art implicit-front tracking schemes, such as the
optimised predictor-corrector scheme of Zia and Lecampion,78 and the algorithms based on a linear extrapolation of 𝐾𝐼 .80
In conclusion, fast and accurate convergence on the fracture length is ensured with the 𝐽𝐻𝐹𝑀 -algorithm independently to
the toughness-viscous regime of propagation. Such a result suggests the 𝐽𝐻𝐹𝑀 -algorithm may be a promising option for
use in reservoir-scale models of hydraulic fracturing.

7.5 Extensions and generalisations

In this work, two novel 𝐽−integral’s have been described for estimating the energy release rate of a hydraulic fracture, and
a novel methodology, the 𝐽𝐻𝐹𝑀 −algorithm, has been derived capable of extracting the location of the fracture front from
the calculated energy release rate. Since some of the derivations are complex, it is worth remarking the applicability of
such contributions in more general, 3-dimensional settings. The 𝐽𝐻𝐹𝑀 −integral uses the one-dimensional, incompressible
lubrication equation in the vicinity of the fracture tip, which is suspected to be equally applicable in 3-dimensional settings.
𝐴
The 𝐽𝐻𝐹𝑀 −integral has the additional requirement that the plane-strain aperture asymptote shown in Equation 18 is appli-
cable in 3−dimensions, which may only be the case sufficiently close to the fracture tip. Furthermore, when considering
propagation through layered elastic media, intersection with a natural fracture, or jumps in the confining stress field, the
𝐴
𝐽𝐻𝐹𝑀 -integral would have to be re-derived if the aperture asymptote used herein is no longer valid. The 𝐽𝐻𝐹𝑀 −algorithm
𝐴
inherits the same requirements of the novel 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 -integrals, along with an assumption regarding the drop in
fluid pressure for an increment in crack extension; a feature calibrated by the parameter 𝜆 in Equation 26. Consequently,
calibration of the 𝐽𝐻𝐹𝑀 -algorithm may be necessary when simulating hydraulic fracturing in 3-dimensions.
A common extension of energy-flux integrals like the 𝐽−integral regards the inclusion of shear stress in their calculation.
Recent theoretical works have derived novel formulations for the energy-release rate and the asymptotics at the fracture tip
in the presence of singular shear tractions at the tip.83–86 The relevance of singular shear tractions in hydraulic fracturing
is debated,87,88 potentially because of the origin of the singularity in shear stress arising from the singularity in fluid
pressure,83 which is a result of the nonphysical, but practical, assumption of zero fluid-lag at the fracture tip. Nonetheless, it
is important to understand how the results herein would change in the presence of significant shear tractions. By including
the shear traction terms within the original 𝐽-integral89 or the 𝐽𝐻𝐹𝑀 −integral, the energy release rate can be appropriately
calculated so long as the numerically-calculated shear stresses are accurate. Such a feature may be problematic if the
shear stress is singular, but could be overcome by supplementing the shear stress (Equation (11) in Piccolroaz .et al89 )
𝐴
with the appropriate asymptotic solutions, in a procedure similar to the 𝐽𝐻𝐹𝑀 -integral. We note that if the tip apertures
𝐴
significantly change from that predicted by the multi-scale asymptote in Equation 18 , then both the 𝐽𝐻𝐹𝑀 -integral, and
the 𝐽𝐻𝐹𝑀 -algorithm would incur errors, but the 𝐽𝐻𝐹𝑀 -integral would not deteriorate in accuracy.

8 CONCLUSION

The following three conclusions can be distinguished from this work:

∙ The 𝐽-integral can theoretically estimate the stress intensity factor throughout the toughness-viscous propagation
regimes, but its accuracy is compromised in practice in the viscosity-dominated regime by the inaccurate solution of
fluid pressure at the fracture tip. This is demonstrated via a finite element analysis showing the 𝐽-integral’s significant
over-estimation of the stress intensity factor in such a regime. Furthermore, the accuracy of the 𝐽-integral significantly
PEZZULLI et al. 2185

deteriorates at coarser resolutions in the viscosity-dominated regime, becoming impractical for reservoir scale simula-
tions.
∙ The errors incurred by the 𝐽-integral can be significantly eliminated via its reformulation to the 𝐽𝐻𝐹𝑀 -integral, and
𝐴
its further enhancement to a 𝐽𝐻𝐹𝑀 -integral. The reformulation 𝐽𝐻𝐹𝑀 relies on the lubrication equations valid at the
fracture tip to reformulate pressure gradients in terms of fracture apertures, introducing the propagation velocity as an
𝐴
additional variable within the integral. In addition to the reformulation, the enhancement 𝐽𝐻𝐹𝑀 uses the asymptotic
𝐴
solution for the aperture profile at the fracture tip. The novel 𝐽𝐻𝐹𝑀 and 𝐽𝐻𝐹𝑀 -integrals displayed significant increases
in accuracy in the viscous-dominated propagation regime, and give identical estimates for the stress intensity factor to
the 𝐽-integral in the toughness-dominated regime.
∙ The energy release rate can be inverted to extract the propagation velocity of the fracture. Such a result is achieved by
using existing and novel asymptotic solutions to express the 𝐽𝐻𝐹𝑀 -integral in semi-analytical form, and solving for the
distance to the fracture tip. The procedure provides a novel energy-based alternative to the aperture-based approaches
of state-of-the-art HFM simulators.18–20,24–26,31,34,78 Such a technique is implemented within a 𝐽𝐻𝐹𝑀 -algorithm that con-
𝐴
verges on the fracture front via iteratively calculating the 𝐽𝐻𝐹𝑀 -integral, inverting the result to predict the location of
the fracture tip, and modifying the time step. The 𝐽𝐻𝐹𝑀 -algorithm is shown to accurately and quickly converge on the
location of the fracture front throughout the toughness and viscous propagation regimes, with small but modest errors
in the toughness regime, and promising results with coarse resolutions of the fracture (ℎ∕𝓁 ∼ 0.2).

Overall, this work has introduced the necessary modifications to use the 𝐽-integral in hydraulic fracture mechanics. The
increase in accuracy of the 𝐽𝐻𝐹𝑀 -integrals, and the front-tracking capabilities of the 𝐽𝐻𝐹𝑀 -algorithm provide a general
methodology to simulate hydraulic fracturing throughout the toughness-viscous regime, with potential in reservoir-scale
applications. Furthermore, the work provides a starting point for the simulation of hydraulic fracturing in materials with
non-linear and/or time-dependent behaviour.

AC K N OW L E D G M E N T S
This research received financial support from SNF grant number 200020_204499, ‘Key problems of heat and mass transfer
in the earth’s crust’. The authors would also like to thank Brice Lecampion and Mohammad Vahab for insightful discus-
sions related to hydraulic fracturing.
Open access funding provided by Eidgenossische Technische Hochschule Zurich.

D A T A AVA I L A B I L I T Y S T A T E M E N T
All data that support the findings of this study are available from the corresponding author upon reasonable request.

ORCID
Edoardo Pezzulli https://orcid.org/0000-0001-9866-1832
Saeed Salimzadeh https://orcid.org/0000-0001-7111-971X

REFERENCES
1. Ague JJ. 4.6 - Fluid flow in the deep crust. In: Holland HD, Turekian KK., eds. Treatise on Geochemistry (Second Edition) Oxford: Elsevier.
2014 (pp. 203–247) http://www.sciencedirect.com/science/article/pii/B9780080959757003065. https://doi.org/10.1016/B978-0-08-095975-7.
00306-5
2. Ingebritsen S, Gleeson T. Crustal permeability. Hydrogeol J. 2017;25(8):2221–2224.
3. Rivalta E, Taisne B, Bunger AP, Katz RF. A review of mechanical models of dike propagation: Schools of thought, results and future
directions. Tectonophysics. 2015;638:1–42.
4. Heinrich C, Candela P. Fluids and Ore Formation in the Earth’s Crust. Elsevier; 2014: 1–28. https://linkinghub.elsevier.com/retrieve/pii/
B9780080959757011013. https://doi.org/10.1016/B978-0-08-095975-7.01101-3
5. Fournier RO. Hydrothermal processes related to movement of fluid from plastic into brittle rock in the magmatic-epithermal environment.
Econ Geol. 1999;94(8):1193–1211.
6. Heinrich CA, Driesner T, Stefánsson A, Seward TM. Magmatic vapor contraction and the transport of gold from the porphyry environment
to epithermal ore deposits. Geology. 2004;32(9):761–764.
7. Baria R, Baumgärtner J, Rummel F, Pine RJ, Sato Y. HDR/HWR reservoirs: concepts, understanding and creation. Geothermics.
1999;28(4):533–552.
8. Olasolo P, Juárez MC, Morales MP, D’Amico S, Liarte IA. Enhanced geothermal systems (EGS): a review. Renewable Sustainable Energy
Rev. 2016;56:133–144.
2186 PEZZULLI et al.

9. Gaucher E, Schoenball M, Heidbach O, et al. Induced seismicity in geothermal reservoirs: A review of forecasting approaches. Renewable
Sustainable Energy Rev. 2015;52:1473–1490.
10. Atkinson GM, Eaton DW, Igonin N. Developments in understanding seismicity triggered by hydraulic fracturing. Nat Rev Earth Environ.
2020;1(5):264–277.
11. Bachu S. Sequestration of CO2 in geological media: criteria and approach for site selection in response to climate change. Energy Convers
Manage. 2000;41(9):953–970.
12. Lecampion B, Bunger A, Zhang X. Numerical methods for hydraulic fracture propagation: a review of recent trends. J Nat Gas Sci Eng.
2018;49:66–83.
13. Detournay E. Propagation regimes of fluid-driven fractures in impermeable rocksregimes of fluid-driven fractures in impermeable rocks.
Int J Geomech. 2004;4(1):35–45.
14. Detournay E. Mechanics of hydraulic fractures. Annu Rev Fluid Mech. 2016;48(1):311–339.
15. Garagash DI, Detournay E. Plane-Strain Propagation of a Fluid-Driven Fracture: Small Toughness Solution. J Appl Mech. 2005;72(6):916–
928.
16. Garagash DI. Plane-strain propagation of a fluid-driven fracture during injection and shut-in: asymptotics of large toughness. Eng Fract
Mech. 2006;73(4):456–481.
17. Garagash D, Detournay E, Adachi J. Multiscale tip asymptotics in hydraulic fracture with leak-off. Int J Fluid Mech. 2011;669:260–297.
18. Lecampion B, Peirce A, Detournay E, et al. The impact of the near-tip logic on the accuracy and convergence rate of hydraulic fracture
simulators compared to reference solutions. Effective Sustainable Hydraulic Fracturing. 2013.
19. Peirce A, Detournay E. An implicit level set method for modeling hydraulically driven fractures. Comput Methods Appl Mech Eng.
2008;197(33):2858–2885.
20. Lecampion B, Desroches J. Simultaneous initiation and growth of multiple radial hydraulic fractures from a horizontal wellbore. J Mech
Phys Solids. 2015;82:235–258.
21. Perkowska M, Wrobel M, Mishuris G. Universal hydrofracturing algorithm for shear-thinning fluids: particle velocity based simulation.
Comput Geotech. 2016;71:310–337.
22. Wrobel M, Mishuris G. Hydraulic fracture revisited: Particle velocity based simulation. Int J Eng Sci. 2015;94:23–58.
23. Dontsov EV, Peirce AP. A non-singular integral equation formulation to analyse multiscale behaviour in semi-infinite hydraulic fractures.
Int J Fluid Mech. 2015;781.
24. Dontsov EV, Peirce AP. A multiscale Implicit Level Set Algorithm (ILSA) to model hydraulic fracture propagation incorporating combined
viscous, toughness, and leak-off asymptotics. Comput Methods Appl Mech Eng. 2017;313:53–84.
25. Dontsov EV, Peirce AP. Implementing a Universal Tip Asymptotic Solution into an Implicit Level Set Algorithm (ILSA) For Multiple Parallel
Hydraulic Fractures. American Rock Mechanics Association; 2016. https://www.onepetro.org/conference-paper/ARMA-2016-268
26. Chen X, Zhao J, Li Y, Yan W, Zhang X. Numerical simulation of simultaneous hydraulic fracture growth within a rock layer: implications
for stimulation of low-permeability reservoirs. J Geophys Res: Solid Earth. 2019;124(12):13227–13249. _eprint: https://agupubs.onlinelibrary.
wiley.com/doi/pdf/10.1029/2019JB017942
27. Chen X, Zhao J, Yan W, Zhang X. Numerical investigation into the simultaneous growth of two closely spaced fluid-driven fracturesinves-
tigation into the simultaneous growth of two closely spaced fluid-driven fractures. SPE J. 2019;24(01):274–289.
28. Zia H, Lecampion B, Zhang W. Impact of the anisotropy of fracture toughness on the propagation of planar 3D hydraulic fracture. Int J
Fract. 2018;211(1):103–123.
29. Chen M, Zhang S, Xu Y, Ma X, Zou Y. A numerical method for simulating planar 3D multi-fracture propagation in multi-stage fracturing
of horizontal wells. Pet Explor Dev. 2020;47(1):171–183.
30. Nikolskiy D, Lecampion B. Modeling of simultaneous propagation of multiple blade-like hydraulic fractures from a horizontal well. Rock
Mech Rock Eng. 2020;53(4):1701–1718.
31. Zia H, Lecampion B. PyFrac: A planar 3D hydraulic fracture simulator. Comput Phys Commun. 2020;255:107368.
32. Dontsov EV, Peirce AP. Modeling planar hydraulic fractures driven by laminar-to-turbulent fluid flow. Int J Solids Struct. 2017;128:73–84.
33. Desroches J, Detournay E, Lenoach B, et al. The crack tip region in hydraulic fracturing. Proc R Soc London, Ser A: Mathematical and
Physical Sciences. 1994;447(1929):39–48.
34. Peirce A. Modeling multi-scale processes in hydraulic fracture propagation using the implicit level set algorithm. Comput Methods Appl
Mech Eng. 2015;283:881–908.
35. Gordeliy E, Peirce A. Implicit level set schemes for modeling hydraulic fractures using the XFEM. Comput Methods Appl Mech Eng.
2013;266:125–143.
36. Gordeliy E, Abbas S, Peirce A. Modeling nonplanar hydraulic fracture propagation using the XFEM: an implicit level-set algorithm and
fracture tip asymptotics. Int J Solids Struct. 2019;159:135–155.
37. Peirce A. Implicit level set algorithms for modelling hydraulic fracture propagation. Philos Trans R Soc A: Mathematical, Physical and
Engineering Sciences. 2016;374(2078):20150423.
38. Li FZ, Shih CF, Needleman A. A comparison of methods for calculating energy release rates. Eng Fract Mech. 1985;21(2):405–421.
39. Lim IL, Johnston IW, Choi SK. Comparison between various displacement-based stress intensity factor computation techniques. Int J
Fract. 1992;58(3):193–210.
40. Nejati M, Paluszny A, Zimmerman RW. On the use of quarter-point tetrahedral finite elements in linear elastic fracture mechanics. Eng
Fract Mech. 2015;144:194–221.
41. Anderson TL. Fracture Mechanics: Fundamentals and Applications. CRC Press 2005:630. https://doi.org/10.1201/9781420058215
PEZZULLI et al. 2187

42. Rice JR. A Path integral and the approximate analysis of strain concentration by notches and cracksindependent integral and the approx-
imate analysis of strain concentration by notches and cracks. J Appl Mech. 1968;35(2):379–386.
43. Kuna M. Finite Elements in Fracture Mechanics. 201 of Solid Mechanics and Its Applications. Springer Netherlands; 2013. http://link.
springer.com/10.1007/978-94-007-6680-8. https://doi.org/10.1007/978-94-007-6680-8
44. Anderson TL. Fracture Mechanics : Fundamentals and Applications. 4th ed. CRC Press. 2017. https://www.taylorfrancis.com/books/
9781315370293. https://doi.org/10.1201/9781315370293
45. Moran B, Shih CF. Crack tip and associated domain integrals from momentum and energy balance. Eng Fract Mech. 1987;27(6):615–642.
46. Moran B, Shih CF. A general treatment of crack tip contour integrals. Int J Fract. 1987;35(4):295–310.
47. Karlsson A, Bäcklund J. J-integral at loaded crack surfaces. Int J Fract. 1978;14(6):R311–R318.
48. Shih CF, Moran B, Nakamura T. Energy release rate along a three-dimensional crack front in a thermally stressed body. Int J Fract.
1986;30(2):79–102.
49. Pereira JP, Duarte CA. The contour integral method for loaded cracks. Commun Numer Methods Eng. 2006;22(5):421–432.
50. Garzon J, Duarte C, Pereira J. Extraction of stress intensity factors for the simulation of 3-D crack growth with the generalized finite
element method. Key Eng Mater. 2013;560:1–36.
51. Nejati M, Paluszny A, Zimmerman RW. A disk-shaped domain integral method for the computation of stress intensity factors using tetra-
hedral meshes. Int J Solids Struct. 2015;69-70:230–251.
52. Song H, Rahman SS. An extended J-integral for evaluating fluid-driven cracks in hydraulic fracturing. J Rock Mech Geotech Eng.
2018;10(5):832–843.
53. Taleghani AD, Olson JE. Analysis of Multistranded Hydraulic Fracture Propagation: An Improved Model for the Interaction Between
Induced and Natural Fractures. In: Society of Petroleum Engineers; 2009. https://www.onepetro.org/conference-paper/SPE-124884-MS.
https://doi.org/10.2118/124884-MS
54. Gupta P, Duarte CA. Simulation of non-planar three-dimensional hydraulic fracture propagation. Int J Numer Anal Methods Geomech.
2014;38(13):1397–1430.
55. Khoei AR, Vahab M, Hirmand M. Modeling the interaction between fluid-driven fracture and natural fault using an enriched-FEM tech-
nique. Int J Fract. 2016;197(1):1–24.
56. Paluszny A, Salimzadeh S, Zimmerman RW. Chapter 1 - Finite-Element Modeling of the Growth and Interaction of Hydraulic Fractures
in Poroelastic Rock Formations. In: Wu YS., ed. Hydraulic Fracture Modeling Gulf Professional Publishing. 2018 (pp. 1–19). http://www.
sciencedirect.com/science/article/pii/B9780128129982000011. https://doi.org/10.1016/B978-0-12-812998-2.00001-1
57. Salimzadeh S, Paluszny A, Zimmerman RW. Three-dimensional poroelastic effects during hydraulic fracturing in permeable rocks. Int J
Solids Struct. 2017;108:153–163.
58. Salimzadeh S, Hagerup ED, Kadeethum T, Nick HM. The effect of stress distribution on the shape and direction of hydraulic fractures in
layered media. Eng Fract Mech. 2019;215:151–163.
59. Salimzadeh S, Zimmerman RW, Khalili N. Gravity hydraulic fracturing: a method to create self-driven fractures. Geophys Res Lett.
2020;47(20):e2020GL087563.
60. Ren G, Younis RM. An integrated numerical model for coupled poro-hydro-mechanics and fracture propagation using embedded meshes.
Comput Methods Appl Mech Eng. 2021;376:113606.
61. Settgast RR, Fu P, Walsh SDC, White JA, Annavarapu C, Ryerson FJ. A fully coupled method for massively parallel simulation of hydrauli-
cally driven fractures in 3-dimensions. Int J Numer Anal Methods Geomech. 2017;41(5):627–653. _eprint: https://onlinelibrary.wiley.com/
doi/pdf/10.1002/nag.2557
62. Fu P, Settgast RR, Hao Y, Morris JP, Ryerson FJ. The influence of hydraulic fracturing on carbon storage performance. J Geophys Res: Solid
Earth. 2017;122(12):9931–9949.
63. Garagash D, Detournay E. The tip region of a fluid-driven fracture in an elastic medium. J Appl Mech. 2000;67(1):183–192.
64. Lecampion B, Detournay E. An implicit algorithm for the propagation of a hydraulic fracture with a fluid lag. Comput Methods Appl Mech
Eng. 2007;196(49):4863–4880.
65. Zienkiewicz OC, Taylor RL, Fox D., eds. The Finite Element Method for Solid and Structural Mechanics. 7th ed. Butterworth-Heinemann;
2014. http://www.sciencedirect.com/science/article/pii/B9781856176347000168. https://doi.org/10.1016/B978-1-85617-634-7.00016-8
66. Detournay E, Peirce A. On the moving boundary conditions for a hydraulic fracture. Int J Eng Sci. 2014;84:147–155.
67. Williams ML. Stress singularities resulting from various boundary conditions in angular corners of plates in extension. J Appl Mech.
1952;19:526–528.
68. Williams ML. On the stress distribution at the base of a stationary crack. J Appl Mech. 1957:6.
69. Acheson DJ. Elementary Fluid Dynamics. Oxford Applied Mathematics and Computing Science SeriesOxford. Oxford University Press;
1990.
70. Matthäi SK, Mezentsev AA, Belayneh M. Finite Element - Node-Centered Finite-Volume Two-Phase-Flow Experiments With Fractured
Rock Represented by Unstructured Hybrid-Element Meshes. SPE Reservoir Evaluation & Engineering. 2007;10(06):740–756.
71. Stüben K. A review of algebraic multigrid. In: Sloan D, Süli E, Vandewalle S., eds. Partial Differential Equations. 7 of Numerical Analysis
2000. Elsevier; 2001: 281–309. http://www.sciencedirect.com/science/article/pii/B9780444506160500129. https://doi.org/10.1016/B978-0-
444-50616-0.50012-9
72. Valko P, Economides MJ. Hydraulic Fracture Mechanics. 28. Wiley; 1995.
73. Adachi J, Siebrits E, Peirce A, Desroches J. Computer simulation of hydraulic fractures. Int J Rock Mech Min Sci. 2007;44(5):739–757.
2188 PEZZULLI et al.

74. Bao JQ, Fathi E, Ameri S. Uniform investigation of hydraulic fracturing propagation regimes in the plane strain model. Int J Numer Anal
Methods Geomech. 2015;39(5):507–523. _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/nag.2320
75. Isida M. Effect of width and length on stress intensity factors of internally cracked plates under various boundary conditions. Int J Fract
Mechanics. 1971;7(3):301–316.
76. Gordeliy E, Peirce A. Coupling schemes for modeling hydraulic fracture propagation using the XFEM. Comput Methods Appl Mech Eng.
2013;253:305–322.
77. Nejati M, Paluszny A, Zimmerman RW. A finite element framework for modeling internal frictional contact in three-dimensional fractured
media using unstructured tetrahedral meshes. Comput Methods Appl Mech Eng. 2016;306:123–150.
78. Zia H, Lecampion B. Explicit versus implicit front advancing schemes for the simulation of hydraulic fracture growth. Int J Numer Anal
Methods Geomech. 2019;0(0).
79. Thomas RN, Paluszny A, Zimmerman RW. Growth of three-dimensional fractures, arrays, and networks in brittle rocks under tension
and compression. Comput Geotech. 2020;121:103447.
80. Shauer N, Duarte CA. Improved algorithms for generalized finite element simulations of three-dimensional hydraulic fracture propaga-
tion. Int J Numer Anal Methods Geomech. 2019;43(18):2707–2742. _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/nag.2977
81. Galassi Mea. GNU Scientific Library Reference Manual; 2018. https://www.gnu.org/software/gsl/
82. Dontsov EV. An approximate solution for a plane strain hydraulic fracture that accounts for fracture toughness, fluid viscosity, and leak-off.
Int J Fract. 2017;205(2):221–237.
83. Wrobel M, Mishuris G, Piccolroaz A. Energy release rate in hydraulic fracture: can we neglect an impact of the hydraulically induced shear
stress?. Int J Eng Sci. 2017;111:28–51.
84. Perkowska M, Piccolroaz A, Wrobel M, Mishuris G. Redirection of a crack driven by viscous fluid. Int J Eng Sci. 2017;121:182–193.
85. Shen W, Zhao YP. Combined effect of pressure and shear stress on penny-shaped fluid-driven cracks. J Appl Mech. 2018;85(3).
86. Piccolroaz A, Peck D, Wrobel M, Mishuris G. Energy release rate, the crack closure integral and admissible singular fields in fracture
mechanics. Int J Eng Sci. 2021;164:103487.
87. Linkov A. Response to the paper by M. Wrobel, G. Mishuris, A. Piccolroaz “Energy release rate in hydraulic fracture: can we neglect
an impact of the hydraulically induced shear stress?” (International Journal of Engineering Science, 2017, 111, 28–51). Int J Eng Sci.
2018;127:217–219.
88. Wrobel M, Mishuris G, Piccolroaz A. On the impact of tangential traction on the crack surfaces induced by fluid in hydraulic fracture:
Response to the letter of A.M. Linkov. Int. J. Eng. Sci. (2018) 127, 220–224. Int J Eng Sci. 2018;127:220–224.
89. Piccolroaz A, Peck D, Wrobel M, Mishuris G. Energy release rate, the crack closure integral and admissible singular fields in fracture
mechanics. Int J Eng Sci. 2021;164:103487.
90. Inc WR. Mathematica, Version 12.0; 2019.

How to cite this article: Pezzulli E, Nejati M, Salimzadeh S, Matthäi SK, Driesner T. An enhanced J-integral for
hydraulic fracture mechanics. Int J Numer Anal Methods Geomech. 2022;46:2163–2190.
https://doi.org/10.1002/nag.3383

APPENDIX A: 𝑱𝑯𝑭𝑴 -INTEGRAL DERIVATIONS

A.1 Contour JHFM -integral


The contour equivalent of the 𝐽𝐻𝐹𝑀 -integral is
( ) 𝑥̂ 0
𝜕𝑢𝑗 [ ]𝑥=
̂ 𝑥̂ 0 1
𝐽𝐻𝐹𝑀 = 𝑊 𝛿1𝑖 − 𝜎𝑖𝑗 𝑚𝑖 𝑑Γ + 𝑤𝑝𝑓 𝑥=0 − 𝜇′ 𝑉 ̂
𝑑𝑥, (A.1)
∫Γ 𝜕𝑥1 ̂ ∫0 𝑤
0

where the terms 𝑥̂ and 𝑥̂ 0 are defined as in Figure 3. To derive the contour form of the velocity-dependent 𝐽-integral only
the second term on the right hand side of Equation 7 needs to be considered

𝜕𝑢2
𝐽𝑓 = 𝑝𝑓 𝑚 𝑑Γ, (A.2)
∫Γ
+ +Γ−
𝜕𝑥1 2

similarly denoted 𝐽𝑓 herein, but not to be confused with the domain version Equation 11. Following Song and Rahmann,52
the 𝐽𝑓 term can be reformulated in terms of the aperture

𝜕𝑤
𝐽𝑓 = − 𝑝𝑓 𝑑Γ (A.3)
∫Γ 𝜕𝑥1
+
PEZZULLI et al. 2189

0
𝜕𝑤
=− 𝑝𝑓 ̂
𝑑𝑥, (A.4)
∫𝑥̂ 𝜕 𝑥̂
0

by acknowledging 𝑤 = 𝑢2+ − 𝑢2− , and that 𝑚2 = −1 when it is the outward pointing normal of Γ+ . In the second step of
Equation A.3, the coordinate 𝑥̂ = −𝑥1 has been used, which points inward from the fracture tip. The form in Equation A.3
can be integrated by parts to give

0 𝜕𝑝𝑓
[ ]𝑥=0
̂
𝐽𝑓 = − 𝑤𝑝𝑓 𝑥= + 𝑤 𝑑𝑥̂ . (A.5)
̂ 𝑥̂ 0 ∫𝑥̂ 𝜕 𝑥̂
0

Applying the boundary condition 𝑤(𝑥̂ = 0) = 0 eliminates the need to evaluate the first term on the right side at the
fracture tip 𝑥̂ = 0, which would otherwise eliminate the benefit of the formulation. Denoting 𝑤0 = 𝑤(𝑥̂ = 𝑥̂ 0 ) and 𝑝𝑓0 =
𝑝𝑓 (𝑥̂ = 𝑥̂ 0 ), and substituting the asymptotic Reynold’s equation Equation 9 into the second term of Equation A.5 gives

0
1
𝐽𝑓 = 𝑤 0 𝑝𝑓0 + 𝜇′ 𝑉 𝑑𝑥̂ . (A.6)
∫𝑥̂ 𝑤
0

Upon substitution of Equation A.6 into Equation 7, the contour 𝐽𝐻𝐹𝑀 -integral shown in Equation A.1 is obtained.

A.2 𝐽-integral for viscosity-dominated regime


𝐴
A simplification of the 𝐽𝐻𝐹𝑀 -integral is provided for the case where only viscous-dominated propagation is of interest,
termed 𝐽𝑉 -integral. For such a case, the viscous asymptote

( )1∕3
𝜇′ 𝑉
𝑤=𝛽 𝑥̂ 2∕3 , (A.7)
𝐸′

can be used instead of the D&P asymptote Equation 18. Consequently, the analytical integration is greatly simplified to
yield
( ) 𝑥̂
𝜕𝑢𝑗 𝜕𝑞 1 0
(𝜇′ 𝑉)2∕3 (𝐸 ′ 𝑥̂ 0 )1∕3
𝐽𝑉 = 𝜎𝑖𝑗 − 𝑊𝛿1𝑖 𝑑𝐴 + 𝑤𝑝𝑓 𝑑𝑥̂ − 𝛼 , (A.8)
∫𝐴 ∗ 𝜕𝑥1 𝜕𝑥𝑖 𝑥̂ 0 − 𝑥̂ 𝑞 ∫𝑥̂ 𝑞 𝛽

where 𝛼 is a constant that depends on

3 [ 4∕3 ]
𝛼= 𝑟𝜃 + 4(1 − 𝑟)𝜃1∕3 + 3𝑟 , (A.9)
4

where 𝜃 = 𝑥̂ 𝑞 ∕𝑥̂ 0 and 𝑟 = 𝑥̂ 0 ∕(𝑥̂ 0 − 𝑥̂ 𝑞 ) as before. When 𝑥̂ 0 = 2𝑥̂ 𝑞 , then 𝑟 = 2, 𝜃 = 0.5 and 𝛼 = 2.7141. The formula Equa-
tion A.8 for the 𝐽𝑉 -integral can be used to extract the energy release rate in the viscosity-dominated regime of propagation
when the plateau 𝑞 function and 𝑞̂ is used as in Equation 14.

APPENDIX B: DERIVATION OF SEMI-ANALYTICAL FLUID PRESSURE ASYMPTOTE VALID


THROUGHOUT THE VISCOUS-TOUGHNESS PROPAGATION REGIME
This section is dedicate to deriving the semi-analytical expression for the fluid pressure Equation 26 valid throughout
toughness-viscous regime of propagation. The result is needed to form a solvable Section 27 for the distance to the fracture
𝐴
tip 𝑥 ∗ , given a numerical estimate for the energy release rate is known (which can be obtained by using the 𝐽𝐻𝐹𝑀 -integral).
A little creativity is needed to obtain a regime-independent asymptotic solution for the fluid pressure. Finding recourse
in the fundamental theorem of calculus, the fluid pressure is reformulated as

𝑥̂ 𝜕𝑝𝑓
̂ = 𝑝𝑓𝑖 (𝑥̂ 0 ) +
𝑝𝑓 (𝑥) 𝑑𝑥̂ , (B.1)
∫𝑥̂ 𝜕 𝑥̂
0
2190 PEZZULLI et al.

where 𝑝𝑓𝑖 is the fluid pressure at a distance 𝑥̂ 0 from the current fracture tip 𝓁𝑛,𝑖 (see Figure 7). While 𝑝𝑓𝑖 is obtained numer-
ically, the second term on the right hand side can be evaluated analytically. By using the velocity equation Equation 9, the
pressure gradient can be eliminated to give

𝑥̂
1
̂ = 𝑝𝑓𝑖 (𝑥̂ 0 ) + 𝜇′ 𝑉
𝑝𝑓 (𝑥) 𝑑𝑥̂ . (B.2)
∫𝑥̂ 𝑤 2
0

upon which substituting the asymptotic solution for the aperture 𝑤 completes the formulation. The numerical fluid pres-
sure 𝑝𝑓𝑖 should be evaluated at the distance 𝑥̂ 0 where they are most likely to be accurate, while not compromising the
validity of the velocity equation Equation 9 and aperture Equation 18. As a result, the distance to the farthest node within
the last ring of elements on the fracture is taken.
The integration of the aperture term in Equation B.2 must subsequently be carried out. The result obtained using the
symbolic calculator Mathematica90 is

[( )1∕3 ( √ )]𝑥̂ 0
3𝜇′ 𝑉𝐸 ′2 𝓁𝑘𝑚 2 2 5 1 𝓁𝑘𝑚
̂ = 𝜆𝑝𝑓𝑖 (𝑥̂ 0 ) + 2 ′2
𝑝𝑓 (𝑥) 2 𝐹1 , , ,− . (B.3)
𝛽 𝐾 𝑥 3 3 3 𝛽3 𝑥
𝑥̂

where 𝑥 must be evaluated at 𝑥̂ and 𝑥̂ 0 . The solution is consistent with the current knowledge of the asymptotic solutions
at the fracture tip. Particularly, the result embodies the 𝑥̂ −1∕3 decay towards the fracture tip which is typical of the viscous
regime.17,33,63 Then, as toughness dominated propagation is approached, the transition lengthscale 𝑙𝑘𝑚 → ∞ and the gaus-
sian hypergeometric function 2 𝐹1 (⋅ , ⋅ ; ⋅ ; ⋅) → 0 and thus the term in the square brackets of eq. B.3 vanishes, therefore
progressively flattening the pressure drop, eventually obtaining a constant pressure profile typical of toughness-dominated
propagation.16,17,63
Since no asymptotic solution exists for the absolute value of pressure the numerical fluid pressure 𝑝𝑓𝑖 must be used.
Such a feature may be due to the fact that the solution for the fluid pressure is global which depends on the entire fracture
geometry and boundary conditions.17,66 Hence, asymptotic solutions for the fluid pressure may only exist up to a constant.
The addition of the ad-hoc convergence parameter 𝜆 is necessary to predict the change in absolute pressure 𝑝𝑓𝑖 given
a change in length of the fracture. Choosing 𝜆 = 1 suggests that fluid pressure changes only as a function of the fracture
velocity and the distance to the tip, with no explicit global dependence on the fracture volume. As mentioned above, such
stationary behaviour is true for the fracture aperture and pressure gradient, but not for the fluid pressure.17,66 In both
toughness and viscosity dominated propagation, the fluid pressure decays to the 1∕3 power with time for a plane-strain
fracture fed by a constant flux. Thus 𝜆 = (𝑥̂ 𝑞 ∕𝑥 ∗ )1∕3 makes a suitable guess for the global drop in pressure due to a change
in time. As the algorithm converges to the correct time-step, 𝑥̂ 𝑞 converges to 𝑥∗ , and 𝜆 converges to 1. It should be noted
that the scheme finds it difficult to converge as the exponent within the 𝜆 term diverges away from 1∕3.

You might also like