You are on page 1of 13

Geophysical Journal International

Geophys. J. Int. (2013) 194, 1514–1526 doi: 10.1093/gji/ggt199


Advance Access publication 2013 June 8

Rock deformation models and fluid leak-off in hydraulic fracturing

Viktoriya M. Yarushina, David Bercovici and Michael L. Oristaglio


Department of Geology and Geophysics, Yale University, PO Box 208109, New Haven, CT 06520-8109, USA. E-mail: viktoriya.yarushina@yale.edu

Accepted 2013 May 13. Received 2013 May 10; in original form 2012 July 31
GJI Marine geosciences and applied geophysics

SUMMARY

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Fluid loss into reservoir rocks during hydraulic fracturing is modelled via a poro-elastoplastic
pressure diffusion equation in which the total compressibility is a sum of fluid, rock and
pore space compressibilities. Inclusion of pore compressibility and porosity-dependent per-
meability in the model leads to a strong pressure dependence of leak-off (i.e. drainage rate).
Dilation of the matrix due to fluid invasion causes higher rates of fluid leak-off. The present
model is appropriate for naturally fractured and tight gas reservoirs as well as for soft and
poorly consolidated formations whose mechanical behaviour departs from simple elastic laws.
Enhancement of the leak-off coefficient by dilation, predicted by the new model, may help
explain the low percentage recovery of fracturing fluid (usually between 5 and 50 per cent) in
shale gas stimulation by hydraulic fracturing.
Key words: Geomechanics; Gas and hydrate systems; Permeability and porosity; Plasticity,
diffusion, and creep; Elasticity and anelasticity.

existence of three different regions in the host rock: a filter cake


I N T RO D U C T I O N
formed by trapping solid additives in the fracturing fluid on the
Hydraulic fracturing is an important technology for enhancing oil fracture face; a filtrate zone affected by the invasion of the frac-
and gas production and recovery. During hydraulic fracturing large turing fluid and the reservoir zone with the original reservoir fluid.
amounts of fluid are injected at high pressure into the reservoir rock Each zone has its own physical and transport properties, and the
in order to expand naturally existing fracture networks or create new width, porosity, permeability and pressure drop in each zone are
fractures. This technology is routinely used to enhance production assumed to be constant. Carter’s theory predicts that the rate of
in already high permeability reservoirs such as poorly consolidated fluid loss (by volume) is inversely proportional to the square root
sandstones, as well as in low permeability formations such as oil of time and depends on the viscosities of fracturing and original
and gas shales and tight gas reservoirs. Production is optimized if reservoir fluids, the porosity and permeability of the formation, the
all of the pumped fluid generates fractures. However, in reality some reservoir fluid compressibility and the pressure difference between
fluid leaks off into the formation through porous flow, thus reducing the original reservoir fluid and the fracturing fluid. While Carter’s
the efficiency of the fracturing job. Accurate prediction of leaked theory is useful as a standard model, a number of extensions of the
volume is important for estimating the induced fracture volume, classical theory have tried to account for more complex effects, such
the evolution of pumping pressures and the total amount of fluid as the non-Newtonian nature of the fracturing fluid with a moving
necessary for fracture treatment (Economides & Nolte 2000). The boundary between filtrate and reservoir zones (Biot et al. 1986;
importance of hydraulic fracturing lies beyond the fields of oil and Yi & Peden 1994; Fan & Economides 1995); radial versus linear
gas exploration. Technologies for CO2 sequestration and extraction direction of fluid leak-off from the fracture (Valko & Economides
of geothermal energy from low permeability rocks can be improved 1999); size distribution of polymer particles in the fracturing fluid
by artificial stimulation of rock permeability, which maximizes flow (Shapiro et al. 2007; Lohne et al. 2010) and mixing of fracturing
rates and storage capacity, and enhances the fluid mineral contact and reservoir fluids in the filtrate zone (Settari 1985).
area for underground mineral carbonation reactions (Zimmermann The classic theory does not accurately predict fluid loss for soft
et al. 2009; Kelemen et al. 2011). and poorly consolidated rocks and naturally fractured reservoirs,
Leak-off of fracturing fluid into the formation is typically mod- where a much stronger dependence of leak-off volume on the effec-
elled using Carter’s equation describing 1-D filtration of a viscous tive pressure rise is observed—a phenomenon known as “pressure-
fluid through a rigid porous rock in the direction perpendicular to dependent leak-off ” (Castillo 1987; Barree 1998; Rollins & Hyden
the fracture face (see Settari 1985; Economides & Nolte 2000 for 1998; Meyer & Jacot 2000; Craig et al. 2005; Ramurthy et al. 2009).
an overview of the classical theory). Classical theory postulates the Studies of pressure-dependent fluid loss have proposed leak-off

1514 
C The Authors 2013. Published by Oxford University Press on behalf of The Royal Astronomical Society.
Rock deformation and hydraulic fracturing 1515

coefficients with exponential (Mukherjee et al. 1991; Barree 1998) Compaction is a fundamental process in which the pore space of
or power-law dependence on the effective pressure rise (Castillo sedimentary rocks decreases in response to an applied load. It may
1987; Meyer & Jacot 2000). Warpinski (1991) considered fissured result from the closure of microcracks, grain rearrangement and
rocks with pressure-dependent permeability and derived a leak-off pore collapse, or in some cases from elastic reversible deformation
coefficient that is a power-law function of the logarithm of stress. of the rock. Inelastic irreversible deformation contributes to com-
Pressure-dependent leak-off also results from the model of Fan & paction processes in sedimentary rocks, especially in weak and frac-
Economides (1995), where the dependence is given implicitly as tured media such as shale, soft limestone, sandstone, coalbeds and
a solution to the coupled fluid flow equations in the invaded and unconsolidated sands (Byerlee 1978; Barton et al. 1985; Franquet &
reservoir zones. These authors considered non-Newtonian fractur- Economides 1999; Connolly & Podladchikov 2000; Gil et al. 2005;
ing fluids and assumed a classical Kozeny–Carman type relationship Yao et al. 2010). Inelastic rate-independent (plastic) deformation
between permeability and porosity. All these models assume, how- leads to the dependence of the elastic moduli on effective pressure
ever, that porosity and total compressibility of the fractured rock do and loading history. These phenomena are observed even at mod-
not change with increasing pressure. The effects of poroelasticity erate effective pressures and are enhanced by the presence of shear
on fracture growth were examined in the works of Detournay et al. stresses. Experimental results of Franquet & Economides (1999)

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


(1990) and Cleary (1979) who concluded that the poroelastic nature on unconsolidated sands show pressure and history dependence of
of the fracture walls leads to a back stress, which causes the matrix Young’s modulus and Poisson’s ratio for effective pressures as low
to dilate in such a way as to reduce the fracture width. These studies as 7 MPa. Poulsen & Abass (1993) report the onset of plasticity
used Carter’s leak-off model. Recently, 3-D numerical models for in poorly consolidated sandstones at effective pressure of 700 kPa.
hydraulic fracturing were developed which couple reservoir flow, Here, we consider three different types of rock rheologies that can
fracture propagation and poroelastic deformation of the reservoir be associated with a deformable porous medium, in particular, lin-
matrix (see Dean & Schmidt 2008; Ji et al. 2009 and references ear poroelastic compaction, non-linear poroelastic compaction due
therein). These models allow porosity or permeability to change to opening/closure of pre-existing microfractures and microdamage
in response to external loading and show that pressure-dependent resulting in irreversible elastoplastic pore collapse.
permeability can greatly increase leak-off. Fluid leak-off is treated Conventional hydraulic fracture models based on linear elastic
implicitly by solving for fluid flow through a poroelastic matrix. fracture mechanics fail for soft and fractured rocks (Valko & Econo-
Use of such complex numerical models can still benefit from better mides 1994). Non-linear models that consider an inelastic process
analytical leak-off models for fluid loss. zone around the fracture tip use a continuum damage approach
Here, we focus on more realistic rheological behaviour and elu- (Valko & Economides 1994) or plasticity theory (Hutchinson 1968;
cidate its effect in a 1-D flow geometry. In particular, we exam- Rice & Sorensen 1978; Dong & Pater 2008; Yao et al. 2010).
ine the effects of elastoplastic compaction/decompaction and two- As a fracture propagates it leaves an area of weakened rock adja-
parameter porosity-dependent permeability on leak-off behaviour cent to the fracture faces. Even moderate fluid pressures would be
during hydraulic fracturing. We suggest that poromechanical be- enough to cause inelastic deformation during the leak-off process
haviour yields higher than conventional leak-off, and thus leads to through the fracture walls. Numerical models of a fracture propagat-
unique features that can be observed in the field data and should be ing through a poroelastic matrix confirm this conclusion showing
accounted for in reservoir modelling. that the critical stress levels necessary to initiate failure (plasticity)
To our knowledge, our model is the first to consider ductile com- of the host rock are reached all around the fracture surface and
paction/decompaction of the formation as a mechanism for fluid not only at the fracture tips as suggested by fracture models for
leak-off in hydraulic fracturing. Even though recently developed non-porous rocks. Microseismic monitoring of hydraulic fractures
numerical reservoir simulators are potentially able to perform this also shows that microseismic events can be interpreted as manifesta-
task, no explicit expressions for the leak-off coefficient for such tions of plastic behaviour along the fracture boundaries (Settari et al.
a deformational mechanism have been given in the literature, and 2002).
no systematic analysis of the influence of reservoir permeability or
rheology on leak-off has been presented previously.

M AT H E M AT I C A L M O D E L

P E R M E A B I L I T Y A N D C O M PA C T I O N O F We consider the 1-D diffusion of fracturing fluid into the reservoir,


P O RO U S RO C K S while neglecting the formation of an external filter cake (Fig. 1). We
treat the reservoir rock as a simple deformable two-phase mixture
Numerous studies indicate that permeability is a dynamic property (Mckenzie 1984; Bercovici & Ricard 2003; Tantserev et al. 2009;
that changes in response to an applied load. Relationships between Yarushina 2009). Flow through the invaded and reservoir regions
permeability and effective pressure (Nathenson 1999; Gutierrez will have similar descriptions, but with different material parameters
et al. 2000; Ghabezloo et al. 2009), or between permeability and and boundary conditions; therefore, we do not distinguish between
porosity (Yang & Aplin 2010; Gamage et al. 2011), exist for many the two regions and treat them in a unified way. The mathematical
sedimentary rocks. Experimental studies on the relationship be- formulation consists of two equations for two unknown functions
tween permeability, porosity and effective pressure suggest that (see Appendix A), porosity, ϕ, and effective pressure, pe , which is
the correlated changes in pressure and porosity cause permeability lithostatic pressure minus fluid pressure:
to change. This implies that compaction is responsible for per-
meability variations in a great variety of porous rocks including dϕ = −ϕcϕ d pe , (1)
sandstones, shale, unconsolidated sands and tight gas sands (David
et al. 1994; Davies & Davies 2001; Dong et al. 2010). Moreover,
 
higher-pressure sensitivity is observed at very low effective pres- d pe ∂ k ∂ pe
ct = . (2)
sures typical for hydraulic fracturing. dt ∂x μ ∂x
1516 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

However, for fractured rocks, porosity can have a stronger de-


pendence on effective pressure. Thus, the effective pore space com-
pressibility will depend both on porosity and effective pressure. The
simplest power law of the form:
ϕ = ϕ0 ( pe / p0 )−q , (6)
was proposed by Dong et al. (2010) for shale and sandstone based
on their experimental data, where ϕ0 is a porosity at a reference
pressure p0 and q is a dimensionless material constant. The sug-
gested values of q are 0.024–0.056 for sandstone and 0.006–0.046
for shale, ϕ0 = 0.18−0.22 and ϕ0 = 0.09−0.15 at p0 = 0.1 MPa.
The corresponding effective compressibility takes the form
cϕ = q/ pe . (7)

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


The non-linear compaction relation (6) reproduces the initial
stages of compaction experiments (as indicated in the data of
Vajdova et al. 2004), which are associated with microcrack clo-
Figure 1. 1-D model of fluid leak-off. Fracturing fluid filtrates into the rock sure; at intermediate effective pressures, however, experimental
perpendicular to the fracture face without forming an external filter cake. compaction data can be reproduced with the linear poroelastic rela-
In the invaded zone it replaces the reservoir fluid pushing it further into tion (5). At high pressures above some critical value that depends on
the rock matrix. The reservoir zone is filled with an original reservoir fluid. porosity and yield strength of intact rock (Carroll 1980), compaction
Fracturing and reservoir fluids each are characterized by viscosities μi , μr is associated with pore collapse and can be described by models
and compressibilities c f,i , c f,r , respectively. The compaction process due that involve inelastic deformation (Zhu et al. 1997; Vajdova et al.
to fluid invasion alters the permeability and porosity in both regions. The 2004).
boundary between these regions x = s(t) moves to the right as more fluid In weak rocks with low cohesion, the propagation of a fracture
enters the rock.
is accompanied by damage due to the creation of microfractures
and frictional sliding of pre-existing fractures (Barree 1998). The
effective compressibility in such rocks is expected to be of the form
Here t is time and x is a space coordinate perpendicular to the frac- (see Appendix B):
ture face. Eqs (1) and (2) include several parameters, namely, fluid 
viscosity μ, permeability k, effective pore space compressibility cϕ β if pe ≥ −Ys
cϕ = , (8)
−q
and total compressibility ct . For permeability, we adopt a power law β p̃ if − Y cotφ ≤ pe < −Ys
frequently used in the literature (Connolly & Podladchikov 2000;
where p̃ = (1 + sinφ)(1 + tanφpe /Y ), q = (1 + cscφ)/(1 + sinψ)
Dong et al. 2010):
and Ys = Y cosφ/(1 + sinφ).
k = k0 (ϕ/ϕ0 )n , (3) Here, β = 1/G s for idealized cylindrical pores, φ is the friction
angle, ψ is the dilation angle and Y is the cohesion. The upper
where k0 , ϕ0 , n are material parameters. limit for pe in eq. (8) defines the onset of elastoplastic deformation
For elastic rocks, the effective pore space compressibility, cϕ , is while the lower limit corresponds to the fully plastic state when all
given by fractures in the rock matrix are activated and the fluid flow regime
cϕ = β, (4) may be better described as a flow through the open fracture network
based, for instance, on a dual porosity model (Li et al. 2007). The
where β is a material parameter that in general depends on tempera- typical value of friction angle for most of the rock types is close
ture, pore geometry and mechanical properties of grains. Available to π/6 at high normal stresses (Byerlee 1978), although deviations
experimental data suggest that β = (0.44 to 3.30) × 10−3 MPa−1 for from this value are not uncommon. Salisbury et al. (1991) report
sandstone and β = (0.14 to 1.58) × 10−3 MPa−1 for shale (David the friction angle to be within the range 0.26–0.35 for different
et al. 1994; Dong et al. 2010). Theoretical predictions vary from shale samples. The same authors give cohesion between 0.3 and
β = 1/G s for idealized cylindrical pores to β = 1/(π εG s ) for 0.55 MPa. The microdamage to the rock during the propagation of
penny-shaped ellipsoidal cracks (Mavko et al. 1998; Yarushina the main fracture may significantly reduce the strength of the rock
2009), where G s is the shear modulus of intact rock and ε is the crack and lead to substantial elastoplastic deformation during the leak-off
aspect ratio. If we assume that typical rock shear modulus is 30 GPa stage. Plastic deformation might increase effective compressibility
then rock with cylindrical pores will have β = 0.03 × 10−3 MPa−1 , of the rock several orders of magnitude with increasing effective
which is much smaller than values cited above. However, the pressure (Fig. 2).
presence of elliptical cracks (aspect ratio 0.01) will significantly The total rock compressibility ct is given by the following equa-
increase compressibility and result in β = 1.06 × 10−3 MPa−1 . tion (Appendix A):
Smaller aspect ratios (more elongated pores) will give even higher
values of pore compressibility. Integration of eqs (1) and (4) ϕcϕ ϕ 2 cs
ct = + + ϕc f , (9)
gives an exponential relationship between porosity and effective 1−ϕ 1−ϕ
pressure: which states that the total compressibility of porous rock filled
with a slightly compressible fluid is a combination of compressibil-
ϕ = ϕ0 exp (−β ( pe − p0 )) , (5)
ity of the pore space (first term), compressibility of solid mineral
which was first empirically obtained by Athy (1930) and widely grains (second term) and fluid compressibility (last term). Using fol-
used since then for a variety of sedimentary rocks. lowing parameters cs = 0.1 GPa−1 , c f = 0.45 GPa−1 , β = 1 GPa−1
Rock deformation and hydraulic fracturing 1517

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Figure 2. Effective compressibility of elastoplastic porous rock given by
eq. (8) as a function of effective pressure. Figure 3. Evolution of transient leak-off coefficient for invaded zone with
time obtained for poroelastic rocks with small porosity by numerical in-
tegration of coupled eqs (1) and (2) for n = 3 (solid line) and material
parameters summarized in Table 2. Dotted line shows the steady-state limit
and ϕ = 10 per cent—which are typical of shale and water as a hy- to the leak-off coefficient that is used in similarity solutions and obtained by
numerical integration of eq. (C10).
drofracturing fluid—and the elastic relation (4) for cϕ , we get that
ct = 0.1111 + 0.0011 + 0.0455. Therefore, the pore space com-
pressibility gives the highest contribution to total compressibility
with the fluid compressibility approximately one order of magni- √
at 0 ≤ ξi < c f,i μi η, and in the reservoir zone one obtains
tude less. If rock is described by one of the inelastic relations (7) or √  √ 
(8), then the contribution from rock compaction will be even more f b − erf μr c̃r η erf cr∗ ξr
fr = √  − ( f b − 1) √ , (12)
pronounced since inelastic deformation increases the compressibil- 1 − erf μr c̃r η 1 − erf μr c̃r η
ity of the rock matrix by several orders of magnitude (Fig. 2). With √
high values of pore compressibility characteristic to inelastic rock at c f,r μr η < ξr < ∞, where
behaviour, the total compressibility would be dominated by the √
compressibility of pore space. Neglecting it in calculations of fluid ξ = x/ at,
leak-off may lead to inaccurate results. We can, however, neglect c̃ β
the compressibility of the solid mineral grains. In this case, the total c∗ = = + 1, and
cf (1 − ϕ0 f ) c f
compressibility becomes
  
−1
ϕcϕ π μi  
ct = + ϕc f = ϕ c̃. (10) f b = exp (β p) 1 + βη exp c̃i μi η2 erf μi c̃i η .
1−ϕ c̃i
(13)

Indices ‘i’ and ‘r’ refer to the invaded and reservoir zones; p
A N A LY T I C A L R E S U LT S is the fluid pressure rise on the fracture
√ face. The interface between
the two zones moves as s(t) = 2η k0 t/ϕ0 , where η is given as a
The solution to eqs (1) and (2), which describe the coupled processes
solution to an algebraic equation:
of fluid flow and reservoir compaction, can be simplified for linear

c̃i μr  

−1
and non-linear elastic rheologies if a similarity transform is applied
(see Appendix C). Use of a similarity solution implies that we βη π 1 − erf μr c̃r η =
are concerned with steady-state fluid infiltration, and we neglect the μi c̃r
initial stages of leak-off. The numerical solution to the full equations
c̃i √   

shows that this approximation does not introduce significant error (exp (β p) − 1) − βη πexp c̃i μi η2 erf μi c̃i η . (14)
(Fig. 3). Corrections accounting for the initial transient stages of μi
fluid leak-off can be accounted for by spurt loss (Economides &
Nolte 2000). Analytical solutions to eqs (1) and (2), accounting for According to this analytical solution, most of the pressure and
a moving boundary between the invaded and reservoir zones can be porosity drop occurs in the invaded region (Fig. 4). However, the
obtained for the special case when the permeability exponent n = presence of the reservoir region accounts for the speed of propa-
1, and the initial reservoir porosity ϕ0 is small (see Appendix C for gation of the fluid invasion front. Without the reservoir region, the
details). Defining normalized porosity f = ϕ/ϕ0 , then the invaded pressure diffusion front is smooth and moves much faster (dashed
zone has line in Fig. 4). Leak-off of the fracturing fluid changes the poros-
√  ity of the formation, and these changes, which might erroneously
erf ci∗ ξi be considered insignificant, lead to a pressure-dependent leak-off
f i = exp (β p) + ( f b − exp (β p)) √ , (11)
erf μi c̃i η coefficient and also vary the rock permeability.
1518 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Figure 4. Porosity and pressure profiles in the invaded and reservoir zones
predicted by the analytical solution (11) and (12) for n = 1. Dashed line
represents solution for invaded zone without reservoir resistance to filtra-
Figure 5. Total leak-off coefficient for linear poroelastic rocks (15) (solid
tion of fracturing fluid. Values of parameters used for this calculation are
line) as a function of pressure rise obtained for small porosity reservoirs
summarized in Table 2.
with linear porosity-dependent permeability (n = 1) and conventional total
leak-off coefficient (dotted line). Parameters used in these calculations are
summarized in Table 2.
The total leak-off coefficient calculated from the analytical solu-
tion is given by

1 k0 c̃i ϕ0
Ct = exp (β p) η̃, (15) N U M E R I C A L R E S U LT S
β π μi
where The analytical similarity solutions were obtained for a special case

   when porosity is small and permeability changes linearly with
 π μi  
η̃ = erf −1 μi c̃i η 1 − 1 + βη exp c̃i μi η2 porosity (n = 1 in eq. 3). Analysis of the general case is com-
c̃i plicated and analytical solutions for the total leak-off coefficient are


−1  no longer available. However, the nominal leak-off coefficients for


× erf μi c̃i η . (16) the reservoir and invaded zones that account for compaction can
be obtained by considering two simplified systems (Economides &
The coefficient depends on the formation permeability, reser- Nolte 2000). The first system concerns the filtration of the frac-
voir and fracturing fluid viscosities and compressibilities, the pore turing fluid in a non-saturated core. The other one considers flow
compressibility, β and the effective pressure rise during hydraulic of a reservoir fluid in a saturated core. In both cases flow is de-
fracturing, p. The dependence on pressure is given by an exponen- scribed by eqs (1) and (2) with the following boundary and initial
tial term and through the parameter η̃, which is an implicit function conditions:
of pressure (Fig. 5). For comparison, the dotted line in Fig. 5 shows
the classical total leak-off coefficient (Settari 1985; Economides & pe = pfrac at x = 0 and t >0
Nolte 2000): pe = p0 as x → ∞ and t >0
2Cr Ci
Ct = , (17) pe = p0 at t = 0
Ci + Ci2 + 4Cr2
where ϕ = ϕ0 at t = 0. (19)

k0 ϕ0 p k0 ϕ0 (β + c f )
Ci = Ci0 = and Cr = Cr0 = p, (18) Eqs (1), (2) with (19) were solved numerically for three different
2μi π μr
types of rock rheologies: linear poroelastic, non-linear poroelastic
are the nominal leak-off coefficients for the invaded and reservoir and elastoplastic. These three types of behaviour are defined by
zones, respectively. Even though the classical model and our new effective pore compressibilities (4), (7) and (8), respectively. Poroe-
model give very similar results in the specific case for n = 1 (Fig. 5), lastic compaction of the formation greatly reduces the nominal
the predictions differ at more realistic values of n > 1, as discussed leak-off coefficient for the invaded zone; while in the reservoir zone,
in the ‘Numerical Results’ section, and this difference is more compaction of the formation leads to higher fluid loss (Fig. 6). Nu-
pronounced with increasing n. Nevertheless, the idealized case of merically obtained leak-off coefficients for linear poroelastic rocks
n = 1 is useful as an analytical benchmark for the numerical models. can be fit by the following analytical expressions:
However, on the whole, the inclusion of compaction in the leak-off
model gives higher values of the leak-off coefficient as well as k0 ϕ0 p
Ci ≈ 0.6ϕ0 exp (β(n + 20) p) · exp (β p),
leak-off volumes. 2μi
Rock deformation and hydraulic fracturing 1519

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Figure 6. Leak-off coefficients for invaded (Ci ) and reservoir zones (Cr )
in linear poroelastic rocks normalized by conventional leak-off coefficients Figure 8. The ratio of leak-off coefficients for elastoplastic rocks calculated
Ci0 and Cr0 given by eq. (18). Parameters used in these calculations are for invaded zone (Ci ) and reservoir zone (Cr ) and conventional leak-off
summarized in Table 2. coefficients (Ci0 , Cr0 ). Parameters used in these calculations are summarized
in Table 2.

In elastoplastic rocks, the leak-off coefficient depends addition-


ally on cohesion and loading history and changes its curvature at the
transition from elastic to plastic regimes (Fig. 8). Frictional plastic-
ity in rocks leads to higher values of leak-off in comparison with
purely elastic rocks at intermediate pressures (Figs 6–8). However,
as pressure increases, the leak-off in elastoplastic rocks reaches its
limiting value and does not change further until full plastic pore
collapse occurs (Fig. 12). For all three types of rocks, an increase in
n leads to an increase in leak-off. In the examples shown, leak-off
coefficients may increase by a factor of 4 as n increases from 10 to
50. Therefore, the leak-off coefficient may be significantly under-
estimated if the reservoir compaction is neglected for rocks with
strong porosity dependence of permeability (i.e. with high values
of n). For example, shales generally have large permeability expo-
nents, in the range n = 24.9−55.5 (Dong et al. 2010). David et al.
(1994) report n = 4.6−25.4 for five different sandstones. Enhance-
ment of the leak-off coefficient by compaction may help explain
Figure 7. The ratio of leak-off coefficients for non-linearly elastic rocks the low percentage recovery of ‘frac fluid’ (usually between 5 and
with effective compressibility defined by eq. (7) calculated for invaded zone 50 per cent) in gas shale stimulation by hydraulic fracturing (King
(Ci ) and reservoir zone (Cr ) and conventional leak-off coefficients Ci0 and 2012).
Cr0 given by eq. (18). Parameters used in these calculations are summarized The total leak-off coefficients may be obtained by simultaneous
in Table 2. numerical solution of the governing equations (eqs 1 and 2) for
the invaded and reservoir regions separated by a moving boundary
(see Appendix C for details). The classical model, (17) and (18),
underestimates the total leak-off coefficient up to a factor of two
and
in linear poroelastic rocks and up to a factor of three in non-linear
n−1 k 0 ϕ0   poroelastic rocks at high values of n (Figs 9 and 10). The error
Cr ≈ 1.1 p β + c f,r exp (βn p) · exp (β p). introduced by the classical model becomes more significant with
n−2 4π μr
an increase in fracturing pressure. The numerically calculated total
Similar results hold for rocks in which compaction is described by leak-off coefficient for poroelastic rocks is fit to analytical expres-
the power law (6) (Fig. 7). In these rocks, compressibility is a func- sions for the case when compressibility of a reservoir fluid is at
tion of pressure. Therefore, for the comparison with the classical least three orders of magnitude smaller than compressibility of a
model, the pore compressibility was approximated by the constant fracturing fluid (e.g. gas as reservoir fluid and water as a fractur-
number ct = (1.04 × 10−3 + c f ) · 0.09, which gives the best fit to ing fluid). For n ≤ 10, reasonable fit is obtained by the classical
the power-law compaction curve. eq. (17), where Ci is corrected for the presence of compaction and
1520 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Figure 11. Comparison of the total leak-off coefficient Ct for poroelastic
Figure 9. The ratio of the total leak-off coefficient for linear and non- rocks obtained by numerical solution (solid line), Carter’s eqs (17) and (18)
linear poroelastic rocks (Ct ) and total conventional leak-off coefficient (Ct0 ) (dotted line) and with use of approximate analytical expressions (20) and
defined by eqs (17) and (18). (21) (knotted line) for two different permeability exponents, n = 8 (left
panel) and n = 30 (right panel).

Comparison of approximate solutions (20) and (21) with numer-


ical calculations shows that the error introduced by the analytical
approximation grows with increasing n and p but still remains
relatively small (Fig. 11). In general, if the leak-off mechanism is
dominated by deformation, then the classical separation of the total
leak-off coefficient into nominal coefficients for the invaded and
reservoir regions (e.g. see eq. 17) leads to erroneous results. For
example, in the case of non-linear elastic rocks, the fit with the
classical equation leads to significant error. In such cases, the to-
tal leak-off coefficient can be obtained only by solution to the full
system involving both regions.

C O M PA R I S O N O F T H E M O D E L T O
Figure 10. Total leak-off coefficient per unit area in linear and non-linear F I E L D D ATA
poroelastic rocks as a function of a pressure drop during hydraulic fracturing. Leak-off due to rock deformation is readily observed when the nor-
Conventional total leak-off coefficient is plotted for reference as a dotted malized leak-off coefficient is plotted versus pressure on a semi-log
line.
plot. All three types of rheology considered herein give their own
distinctive features. Both linear elastic and elastoplastic deforma-
tion give leak-off versus pressure relations with negative curvature
so that d2 ln C/d p2 ≤ 0. However, the relation with elastoplas-
porosity-dependent permeability in the following manner:
ticity flattens at higher pressures leading to d2 ln C/d p2 = 0 just

k0 ϕ0 p (1 + exp (nβ p)) (1 + exp (β p)) before full pore collapse leading to rock failure (Fig. 12). Non-
Ci = . (20) linear elastic deformation, as described by (6) and associated with
8μi
microcrack opening, yields d2 ln C/d p2 ≤ 0 at small pressures
In gas reservoirs most of the porosity increase occurs in the and d2 ln C/d p2 ≥ 0 at moderate to high pressures. The initial
invaded zone, thus the correction to Cr is not needed. For n > 10, stages characterized by negative curvature and associated with rapid
 growth of ln(C/C0 ) (Fig. 13) are due to a jump from almost un-
k0 ϕ0 p (1 + exp (β p)) (1 + exp (nβ p)) detectable leak-off at small pressures (C = 9.25 × 10−9 ms−0.5 at
Ci =
16μi 2 p = 0.1 MPa for n = 30) to one that is orders of magnitude
larger at later stages of deformation (C = 3.83 × 10−6 ms−0.5 at

p = 15 MPa). At intermediate pressures d2 ln C/d p2 ≈ 0, in
(1 + exp (β p))n+1
+ ⎠. (21) which case the leak-off curves can be approximated as inclined
2n straight lines.
Rock deformation and hydraulic fracturing 1521

sure relation changes its curvature, presumably corresponding to a


change from opening of pre-existing microfractures to poroelasto-
plastic deformation; poroelastic deformation results in the peak
above the straight line that flattens towards the pressure maximum
(Fig. 14a). While the linear segment could potentially be explained
based on the model of Warpinski (1991) for naturally fractured
reservoirs, the highly non-linear upper segment cannot be repro-
duced using the model of Warpinski (1991). Nor can it be explained
with the linear elastic model, which yields much smaller values
of leak-off coefficient. Another field example is from an injec-
tion/falloff test in the Main Almond formation in SW Wyoming
(Fig. 14b). These leak-off data show the pressure-independent por-
tion of the leak-off ratio plot below 4000 psi, followed by rapid
growth of the leak-off coefficient, characteristic of linear poroelastic

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


deformation.

Figure 12. Leak-off ratio plot for linear elastic and elastoplastic rocks for
n = 5 and Y = 3.46 MPa (other parameters used in these calculations are C O N C LU S I O N S
summarized in Table 2). Both linear elastic and elastoplastic deformation
lead to concave down curves but the curve for elastoplastic material flattens
When fluid loss into reservoir rock during hydraulic fracturing is
at high stress just before full pore collapse. modelled via pressure diffusion coupled with porosity evolution,
the pressure drop occurs in the narrow invaded zone with dynamic
characteristics such as time-varying thickness, permeability and
compressibility. Inclusion of pore compressibility and porosity-
dependent permeability into the flow model leads to pressure-
dependent leak-off that is more pronounced than in the classical
model. The inferred plasticity (irrecoverable deformation) also al-
lows an estimation of the permeability increase caused by the frac-
turing process. The ability of the rock to dilate in response to the
fluid invasion increases the leak-off rate.
Of the three different types of rock rheologies considered—
namely, linear poroelastic, non-linear poroelastic and poroelasto-
plastic—power-law or frictional elastoplasticity give a better
description of naturally fractured and soft rocks. Poroelastic be-
haviour, on the other hand, may be expected to give the best de-
scription of conventional reservoir. Irrecoverable time-independent
(plastic) deformation gives the highest values of leak-off. Whereas
Carter’s classical model can accurately predict leak-off in poroelas-
tic rocks with low porosity and near-constant permeability (depend-
ing weakly on porosity). New analytical and numerical solutions
Figure 13. Leak-off ratio plot for non-linear elastic rocks. At small pres- show that leak-off in rocks with more strongly porosity-dependent
sures leak-off is orders of magnitude smaller than at higher pressures and permeability or more complicated rheological behaviour will dif-
can be neglected. Note the concave upward segments at higher pressures. fer significantly from the classical theory. The inclusion of rock
Parameters used in these calculations are summarized in Table 2. compaction and decompaction may be important for describing
naturally fractured and tight gas reservoirs as well as CO2 injection
sites and enhanced geothermal systems; it may also be important to
Field examples (Mukherjee et al. 1991; Barree 1998; John- model soft and poorly consolidated formations whose mechanical
son et al. 1998) demonstrate that all three types of deforma- behaviour departs from the simple elastic laws and/or have strongly
tional mechanisms can occur (Fig. 14). Here, we consider sur- porosity-dependent permeability.
face pressure data from pre-fracturing injection tests in the Upper
Lewis sand, Mesa Verde group, Wyoming (Fig. 14a), which can
be characterized as a tight reservoir. These data exhibit pressure-
independent leak-off at low surface pressure (up to 11 MPa) in
which the deformational mechanisms give negligible leak-off. As AC K N OW L E D G E M E N T S
pressure increases, non-linear elastic deformation due to opening Authors are grateful to Nina Simon for helpful discussions. This
of pre-existing cracks (solid red line in Fig. 14a) gives good ap- work was supported in part by a grant (DE-FE0004375) from
proximation to a nearly linear segment on the leak-off curve with the National Energy Technology Laboratory of US Department of
d2 ln C/d p2 ≈ 0. At yet higher pressures, the leak-off versus pres- Energy.
1522 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Figure 14. Field examples presenting characteristic features of leak-off due to elastic (a) and non-linear elastic and elastoplastic (b) deformation of porous
rocks (based on Barree 1998). (a) The first part of the curve was fitted by non-linear elastic deformation mechanism with the following parameters: p0 =
4.3 MPa, q = 0.05, n = 7; the upper part of the curve was reproduced using elastoplastic model with β = 6 × 10−3 MPa−1 , n = 7, c = 3.46 MPa, p0 =
4.1 MPa (solid red line). The red dotted line represents the best fit by linear elastic model for β = 4.7 × 10−2 MPa−1 , n = 5, p0 = 4 MPa. Here, p0 is the
effective pressure corresponding to the onset of the leak-off due to deformational mechanism. (b) The red curve shows theoretical prediction based on linear
poroelastic model with β = 7 × 10−2 MPa−1 , n = 10, p0 = 0.6 MPa.

1. Nomenclature.
Symbol Meaning Units
c f , cs , ct Compressibilities of fluid and solid, total compressibility MPa−1
Ci , Cr , Ct Leak-off coefficients for the invaded and reservoir zones, total leak-off coefficient ms−1/2
Ci0 , Cr0 , Ct0 Carter’s leak-off coefficients ms−1/2
cϕ Pore compressibility MPa−1
g Gravity ms−2
Gs Solid shear modulus MPa
k0 , k Formation permeability before and during hydraulic fracturing m2
n Porosity exponent for k
p f , pt Fluid pressure, total (lithostatic) pressure MPa
pe Effective pressure, pe = pt − p f MPa
p0 Initial formation effective pressure MPa
p Pressure drop during hydraulic fracturing MPa
q Porosity exponent for effective pressure
t Time s
v f , vs Fluid and solid velocity ms−1
x Distance from the fracture face m
Y Cohesion MPa
β Coefficient of pore compressibility MPa−1
μ Fluid viscosity Pas
ξ Similarity variable, eq. (C2)
ρ f , ρs Fluid and solid density kgm−3
ϕ0 , ϕ Formation porosity before and during hydraulic fracturing
φ Friction angle
ψ Dilation angle

2. Simulation input.
REFERENCES
Parameter Value used
Athy, L.F., 1930. Density, porosity and compaction of sedimentary rocks,
Initial formation porosity 0.1 Bull. Am. Assoc. Pet. Geol., 14, 1–24.
Initial formation permeability, m2 1 × 10−19 Barree, R.D., 1998. Applications of pre-frac injection/falloff tests in fis-
Elastic pore compressibility β, 1/MPa 5 × 10−3 sured reservoirs—field examples, paper SPE 39932, in SPE Rocky Moun-
Fracturing fluid viscosity, Pa s 1 × 10−3 tain Regional/Low-Permeability Reservoirs Symposium and Exhibition,
Reservoir fluid viscosity, Pa s 1 × 10−6 Denver, CO, USA, pp. 277–288.
Fracturing fluid compressibility, 1/MPa 0.45 × 10−3 Barton, N., Bandis, S. & Bakhtar, K., 1985. Strength, deformation and
Reservoir fluid compressibility, 1/MPa 10 conductivity coupling of rock joints, Int. J. Rock Mech. Min., 22, 121–
Compaction power law exponent q 0.032 140.
Friction angle φ, rad π/6 Bercovici, D. & Ricard, Y., 2003. Energetics of a two-phase model of litho-
Dilation angle ψ, rad 0 spheric damage, shear localization and plate-boundary formation, Geo-
Cohesion, MPa 7.5 phys. J. Int., 152, 581–596.
Rock deformation and hydraulic fracturing 1523

Biot, M.A., Masse, L. & Medlin, W.L., 1986. A two-dimensional theory of Jaeger, J.C., Cook, N.G.W. & Zimmerman, R.W., 2007. Fundamentals of
fracture propagation, SPE Prod. Eng., 1, 17–30. Rock Mechanics, 4th edn, Blackwell Pub., Malden, MA.
Byerlee, J., 1978. Friction of rocks, Pure appl. Geophys., 116, 615–626. Ji, L.J., Settari, A. & Sullivan, R.B., 2009. A novel hydraulic fracturing
Carroll, M.M., 1980. Compaction of dry or fluid-filled porous materials, J. model fully coupled with geomechanics and reservoir simulation, SPE J.,
Eng. Mech. Div.-ASCE, 106, 969–990. 14, 423–430.
Carter, J.P., Booker, J.R. & Yeung, S.K., 1986. Cavity expansion in cohesive Johnson, R.L., Dunn, K.P., Bastian, P.A., Hopkins, C.W. & Conway, M.W.,
frictional soils, Geotechnique, 36, 349–358. 1998. Qualifying hydraulic fracturing effectiveness in tight, naturally frac-
Castillo, J.L., 1987. Modified fracture pressure decline analysis including tured reservoirs by combining three-dimensional fracturing and reservoir
pressure-dependent leakoff, paper SPE 16417, in SPE/DOE Low Perme- simulators, paper SPE 49048, in SPE Annual Technical Conference and
ability Symposium, Denver, CO, USA, pp. 273–281. Exhibition, New Orleans, LA, USA, pp. 1–23.
Cleary, M.P., 1979. Rate and structure sensitivity in hydraulic fracturing Kelemen, P.B., Matter, J., Streit, E.E., Rudge, J.F., Curry, W.B. & Blusztajn,
of fluid-saturated porous formations, paper SPE 79–0127, in 20th U.S. J., 2011. Rates and mechanisms of mineral carbonation in peridotite: nat-
Symposium on Rock Mechanics, American Rock Mechanics Association, ural processes and recipes for enhanced, in situ CO2 capture and storage,
Austin, TX, pp. 127–142. Annu. Rev. Earth planet. Sci., 39, 545–576.
Connolly, J.A.D. & Podladchikov, Y.Y., 2000. Temperature-dependent vis- King, G., 2012. Hydraulic fracturing 101, paper SPE 152596, in SPE Hy-
coelastic compaction and compartmentalization in sedimentary basins, draulic Fracturing Technology Conference, Society of Petroleum Engi-

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Tectonophysics, 324, 137–168. neers, The Woodlands, TX, USA, pp. 1–80.
Craig, D.P., Eberhard, M.J., Ramurthy, M., Odegard, C.E. & Mullen, R., Li, Y., Guo, J., Zhao, J. & Yue, Y., 2007. A new model of fluid leak-off in
2005. Permeability, pore pressure, and leakoff-type distributions in Rocky naturally fractured gas fields and its effects on fracture geometry, J. Can.
Mountain basins, SPE Prod. Facil., 20, 48–59. Petrol. Technol., 46, 12–16.
David, C., Wong, T.F., Zhu, W.L. & Zhang, J.X., 1994. Laboratory mea- Lohne, A. et al., 2010. Formation-damage and well-productivity simulation,
surement of compaction-induced permeability change in porous rocks— SPE J., 15, 751–769.
implications for the generation and maintenance of pore pressure excess Mavko, G., Mukerji, T. & Dvorkin, J., 1998. The Rock Physics Handbook :
in the crust, Pure appl. Geophys., 143, 425–456. Tools for Seismic Analysis in Porous Media, Cambridge University Press,
Davies, J.P. & Davies, D.K., 2001. Stress-dependent permeability: charac- Cambridge, New York.
terization and modeling, SPE J., 6, 224–235. Mckenzie, D., 1984. The generation and compaction of partially molten
Dean, R.H. & Schmidt, J.H., 2008. Hydraulic fracture predictions with a fully rock, J. Petrol., 25, 713–765.
coupled geomechanical reservoir simulator, in SPE Annual Technical Meyer, B.R. & Jacot, R.H., 2000. Implementation of fracture calibration
Conference and Exhibition, paper SPE 116470, Society of Petroleum equations for pressure dependent leakoff, paper SPE 62545, in SPE/AAPG
Engineers, Denver, CO, USA, pp. 1–12. Western Regional Meeting, Long Beach, CA, USA, pp. 1–20.
Detournay, E., Cheng, A.H.D. & Mclennan, J.D., 1990. A poroelastic PKN Mukherjee, H., Larkin, S. & Kordziel, W., 1991. Extension of fracture
hydraulic fracture model based on an explicit moving mesh algorithm, pressure decline curve analysis to fissured formations, paper SPE 21872,
J. Energ. Resour.-ASME, 112, 224–230. in Rocky Mountain Regional Metteng and Low-Permeability Reservoirs
Dong, J.J., Hsu, J.Y., Wu, W.J., Shimamoto, T., Hung, J.H., Yeh, E.C., Wu, Symposium, Denver, CO, USA, pp. 671–685.
Y.H. & Sone, H., 2010. Stress-dependence of the permeability and poros- Nathenson, M., 1999. The dependence of permeability on effective stress
ity of sandstone and shale from TCDP Hole-A, Int. J. Rock Mech. Min., from flow tests at hot dry rock reservoirs at Rosemanowes (Cornwall) and
47, 1141–1157. Fenton Hill (New Mexico), Geothermics, 28, 315–340.
Dong, Y. & Pater, C.J., 2008. Observation and modeling of the hydraulic Poulsen, D.K. & Abass, H.H., 1993. Hydraulic fracture modeling in for-
fracture tip in sand, paper ARMA08–377, in 42nd US Rock Mechan- mations exhibiting stress-dependent mechanical properties, paper SPE
ics Symposium and 2nd U.S.-Canada Rock Mechanics Symposium, San 26599, in 68th Annual Technical Conference and Exhibition of the Soci-
Francisco, CA, USA, pp. 1–12. ety of Petroleum Engineers, Houston, TX, USA, pp. 881–889.
Economides, M.J. & Nolte, K.G., 2000. Reservoir Stimulation, 3rd edn, Ramurthy, M., Lyons, B., Hendrickson, R.B., Barree, R.D. & Magill, D.R.,
Wiley, Chichester, England. 2009. Effects of high pressure-dependent leakoff and high process-zone
Fan, Y. & Economides, M.J., 1995. Fracturing fluid leakoff and net pres- stress in coal-stimulation treatments, SPE Prod. Oper., 24, 407–414.
sure behavior in Frac&pack stimulation, paper SPE 29988, in Inter- Rice, J.R. & Sorensen, E.P., 1978. Continuing crack-tip deformation and
national Meeting on Petroleum Engineering, Beijing, China, pp. 357– fracture for plane-strain crack growth in elastic-plastic solids, J. Mech.
369. Phys. Solids, 26, 163–186.
Fortin, J., Gueguen, Y. & Schubnel, A., 2007. Effects of pore collapse and Rollins, K. & Hyden, R.E., 1998. Pressure-dependent leakoff in fracturing—
grain crushing on ultrasonic velocities and Vp /Vs , J. geophys. Res., 112, field examples from the Haynesville Sand, paper SPE 39953, in SPE
B08207, doi:10.1029/2005JB004005. Rocky Mountain Regional Meeting/Low Permeability Reservoirs Sympo-
Franquet, J.A. & Economides, M.J., 1999. Effect of stress and stress path on sium and Exhibition, Denver, CO, USA, pp. 433–441.
Young’s modulus and Poisson ratio of unconsolidated rocks: a new idea Salisbury, D.P., Ramos, G.G. & Wilton, B.S., 1991. Wellbore instability
for hydraulic fracturing, in SPE Latin American and Caribbean Petroleum of shales using a downhole simulation test cell, in Rock Mechanics as
Engineering Conference, Caracas, Venezuela, pp. 1–7. a Multidisciplinary Science: Proceedings of the 32nd US Symposium,
Gamage, K., Screaton, E., Bekins, B. & Aiello, I., 2011. Permeability- Norman, OK, USA, pp. 1015–1024.
porosity relationships of subduction zone sediments, Mar. Geol., 279, Settari, A., 1985. A new general model of fluid loss in hydraulic fracturing,
19–36. Soc. Petrol. Eng. J., 25, 491–501.
Ghabezloo, S., Sulem, J., Guedon, S. & Martineau, F., 2009. Effective stress Settari, A., Sullivan, R.B., Walters, D.A. & Wawrzynek, P.A., 2002. 3-D
law for the permeability of a limestone, Int. J. Rock. Mech. Min., 46, analysis and prediction of microseismicity in fracturing by coupled ge-
297–306. omechanical modeling, paper SPE 75714, in SPE Gas Technology Sym-
Gil, I.R., Hart, R., Roegiers, J.C. & Shimizu, Y., 2005. Considerations on posium, Calgary, Alberta, Canada, pp. 1–8.
hydraulic fracturing of unconsolidated formations, in ISRM International Shapiro, A.A., Bedrikovetsky, P.G., Santos, A. & Medvedev, O.O., 2007. A
Symposium–EUROCK 2005: Impact of Human Activity on the Geological stochastic model for filtration of particulate suspensions with incomplete
Environment, Brno, Czech Republic, pp. 155–161. pore plugging, Transp. Porous Med., 67, 135–164.
Gutierrez, M., Oino, L.E. & Nygard, R., 2000. Stress-dependent permeabil- Tantserev, E., Galerne, C.Y. & Podladchikov, Y.Y., 2009. Multiphase flow
ity of a de-mineralised fracture in shale, Mar. Petrol. Geol., 17, 895–907. in multi-component porous visco-elastic media, in 4th Biot Conference
Hutchinson, J.W., 1968. Singular behaviour at the end of a tensile crack in a on Poromechanics, pp. 959–964, eds Ling, H.I., Smyth, A. & Betti, R.,
hardening material, J. Mech. Phys. Solids, 16, 13–31. DEStech, Lancaster, PA.
1524 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

Vajdova, V., Baud, P. & Wong, T.F., 2004. Compaction, dilatancy, and where
failure in porous carbonate rocks, J. geophys. Res., 109, B05204, d ∂ ∂
doi:10.1029/2003JB002508. = + vs .
Valko, P. & Economides, M.J., 1994. Propagation of hydraulically induced
dt ∂t ∂x
fractures—a continuum damage mechanics approach, Int. J. Rock Mech. Substitution of Darcy’s law, eq. (A2) and rheological eqs (A3) and
Min. Sci. Geomech. Abstr., 31, 221–229. (A4) into eq. (A5) gives
Valko, P.P. & Economides, M.J., 1999. Fluid-leakoff delineation in high-  
ρf d pe ρ f cs d p f ϕρ f cs d pt
permeability fracturing, SPE Prod. Facil., 14, 117–130. − ϕcϕ + ϕρ f c f + ϕ 2 −
Warpinski, N.R., 1991. Hydraulic fracturing in tight, fissured media, J. 1−ϕ dt 1−ϕ dt 1 − ϕ dt
Petrol. Tech., 43, 146–209.  
∂ ρ f k ∂p f
Yang, Y.L. & Aplin, A.C., 2010. A permeability-porosity relationship for − = 0. (A6)
mudstones, Mar. Petrol. Geol., 27, 1692–1697. ∂x μ ∂x
Yao, Y., Gosavi, S.V., Searles, K.H. & Ellison, T.K., 2010. Cohesive fracture We approximate the total pressure with lithostatic stress pt =
mechanics based analysis to model ductile rock fracture, paper ARMA10–
−ρs gz, where z is the depth. Hence, d pt /dt = 0 and d p f = −d pe .
140, in 44th US Rock Mechanics Symposium and 5th U.S.-Canada Rock
With this in mind, we can rewrite the hydraulic equation for slightly
Mechanics Symposium, Salt Lake City, UT, USA, pp. 1–8.

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


Yarushina, V.M., 2009. (De)compaction waves in porous viscoelastoplastic compressible fluid for which density can be approximated as a linear
media, PhD thesis, University of Oslo, Oslo. function of fluid pressure:
Yarushina, V.M. & Podladchikov, Y.Y., 2010. Plastic yielding as a frequency  
ρ f − ρ 0f = c f p f − p0f , (A7)
and amplitude independent mechanism of seismic wave attenuation, Geo-
physics, 75, N51–N63. in its final form
Yi, T.C. & Peden, J.M., 1994. A comprehensive model of fluid loss in  
d pe ∂ k ∂ pe
hydraulic fracturing, SPE Prod. Facil., 9, 267–272. ct = , (A8)
Yu, H.-S., 2006. Plasticity and Geotechnics, Springer-Verlag, New York, dt ∂x μ ∂x
NY. where
Zhu, W.L., Montesi, L.G.J. & Wong, T.F., 1997. Shear-enhanced compaction
ϕcϕ ϕ 2 cs
and permeability reduction: triaxial extension tests on porous sandstone, ct = + + ϕc f . (A9)
Mech. Mater., 25, 199–214. 1−ϕ 1−ϕ
Zimmerman, R.W., 1991. Compressibility of Sandstones, Elsevier, Amster- Eq. (A8) must be solved simultaneously with the compaction
dam. eq. (A4) which gives porosity changes due to variations in effective
Zimmermann, G., Tischner, T., Legarth, B. & Huenges, E., 2009. Pressure-
pressure.
dependent production efficiency of an enhanced geothermal system
(EGS): stimulation results and implications for hydraulic fracture treat-
ments, Pure appl. Geophys., 166, 1089–1106.
APPENDIX B: EFFECTIVE
COMPRESSIBILITY IN
P O RO E L A S T O P L A S T I C RO C K S
A P P E N D I X A : D E R I VAT I O N O F
C O U P L E D H Y D R AU L I C A N D Here, we follow the approach of Yarushina & Podladchikov (2010)
C O M PA C T I O N E Q UAT I O N S and Yarushina (2009) and determine cϕ by considering the defor-
mation of a small representative volume element with pores of
In our derivation of the hydraulic eqs (1) and (2), we start from the simplified cylindrical geometry (Fig. B1). Each pore of radius a is
conservation of mass equations for fluid and solid (see Table 1 for subjected to an increasing fluid pressure, p f , and a total pressure,
notations): pt , at the remote boundary. Plane strain conditions are assumed to
∂(ρ f ϕ) ∂(ρ f ϕv f ) ∂(ρs (1 − ϕ)) ∂(ρs (1 − ϕ)vs ) hold throughout the shell. During hydraulic fracturing very high
+ = 0, + = 0. pressures are imposed on the rock. Prior to reaching the rupture,
∂t ∂x ∂t ∂x
rocks might meet a failure criterion and then plastic deformation
(A1)

Darcy’s law of fluid filtration through deformable porous rock


(Mckenzie 1984),
  k ∂p f
ϕ v f − vs = − , (A2)
μ ∂x
poroelastic equations of state for bulk deformation of matrix grains
and fluid (Zimmerman 1991; Yarushina 2009),

cs (d pt − ϕd p f ) = (1 − ϕ) dρs /ρs , c f d p f = dρ f /ρ f , (A3)

and pore space

dϕ = −ϕcϕ d pe , (A4)

where pe = pt − p f is the effective pressure. Summation and sub-


sequent rearrangement of eq. (A1) gives
ρ f dϕ dρ f ρ f dρs ∂   
+ϕ −ϕ + ρ f ϕ v f − vs = 0, (A5)
1 − ϕ dt dt ρs dt ∂x
Figure B1. Cylindrical elastoplastic model of porous rock.
Rock deformation and hydraulic fracturing 1525

will develop around structural heterogeneities in the rock matrix. The similarity transformation
We define irreversible time-independent deformation of rock that √
occurs after stresses reach the failure criterion as plastic irrespec- ξ = x/ at, (C2)
tive of the micromechanism of such deformation, which is often the  
where a = 4k0 / c f μϕ0 , can be applied to eq. (C1) resulting in
interplay of several micromechanical processes, including microc-  
racking, grain crushing, pore collapse and frictional sliding on an c̃ dϕ d ϕ0 c f k dϕ
−2ξ = , (C3)
array of fissures (Vajdova et al. 2004; Fortin et al. 2007). This be- cϕ dξ dξ ϕ cϕ k0 dξ
haviour is characterized by path-dependence and hysteresis. We use
simple and widely used Mohr–Coulomb criterion, which describes where we have made use of eq. (10). Introducing the normalized
not only failure of intact rock that takes place along a plane due to porosity
a shear stress in that plane, as in a fault, but also describes yielding f = ϕ/ϕ0 , (C4)
in rocks and soils (Yu 2006; Jaeger et al. 2007):
and substituting eq. (3) for permeability into (C3) yields
κσθ − σr = Y ∗ + (1 − κ) p f , (B1)
 
c̃ d f d c f n−1 d f
where κ and Y ∗ are the following functions of cohesion, Y, and

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


−2ξ = f . (C5)
friction angle, φ: cϕ dξ dξ cϕ dξ

κ = (1 + sinφ) / (1 − sinφ) , Next, we follow the approach of Fan & Economides (1995) and
(B2) consider simultaneously the invaded and reservoir regions separated

Y = 2Y cosφ/ (1 − sinφ) . by a moving boundary, as fracturing fluid progresses into the forma-
In response to an increase in fluid pressure, the failure region spreads tion. The materials in both regions have the same functional form of
into the shell. Outside the failure region, material is still in the permeability (3) and pore space compressibility. However, the fluid
elastic state. Due to symmetry, the elastoplastic boundary will be a viscosities and compressibilities are different, which influences the
cylindrical surface with radius c. The solution to the force balance pressure and porosity evolution in each region. In the following
equation and specific elastic or plastic constitutive laws yield the we use subscripts ‘i’ and ‘r’ to distinguish the porosity, pressure,
stresses and velocities within the elastic and plastic parts of the shell permeability, fluid viscosity and compressibility in the invaded and
(Carter et al. 1986). reservoir zones. In each of the two regions, (C5) must be satisfied.
Further, we assume that shell is composed of an incompressible The initial and boundary conditions are:
solid material based on the scaling argument that the compressibility
of solid mineral grains is negligible relative to total compressibility. pei = pfrac at x = 0 and t > 0, (C6)
The porosity equation is obtained from the equation:
per = p0 at t = 0 and x > s(t), (C7)
dϕ dV p
= ,
ϕ Vp where x = s(t) is the moving interface between the regions. At the
interface, continuity of pressures and flow velocities is assumed,
where V p is the pore volume. For a cylindrical void, V p = πa , 2
which leads to two more conditions
therefore,
1 dϕ vr  pei = per at x = s(t), (C8)
=2  .
ϕ dt r r=a
ki ∂ pei kr ∂ per ds
Substitution of the plastic velocity vr into the last equation gives = = at x = s(t). (C9)
ϕi μi ∂ x ϕr μr ∂ x dt
the effective compressibility in elastoplastic rocks in the form
Recasting boundary and initial conditions in terms of the new
cϕ = β p̃ −q , (8)
similarity variables must be done separately for each type of rock
where p̃ = (1 + sinφ)(1 + tanφpe /Y ), q = (1 + cscφ)/(1 + sinψ) rheologies.
and Ys = Y cosφ/(1 + sinφ).
This equation holds only if −Y cotφ ≤ pe < −Ys . The upper limit
corresponds to the plasticity onset and the lower limit gives pressure C1: Linearly elastic rocks
for the onset of fully plastic pore collapse when the failure region
occupies the entire body. Physically, this limiting state is unstable First, we consider purely linear elastic rocks in which the effective
and cannot be described by the same type of model. compressibility is given by (4). In this case (C5) reduces to
 
df d df
− 2ξ c∗ = f n−1 , (C10)
APPENDIX C: SIMILARITY dξ dξ dξ
T R A N S F O R M AT I O N T O H Y D R AU L I C
where
E Q UAT I O N C O U P L E D W I T H
D E F O R M AT I O N c̃ β
c∗ = = + 1. (C11)
cf (1 − ϕ0 f ) c f
Eqs (1) and (2) describe fluid flow in a deforming porous rock,
and constitute a system of two equations for two unknowns. For Using (5) and (C4) conditions (C6)–(C9) can be rewritten as
convenience, we can eliminate the effective pressure in eq. (2) using
f i = exp (β p) at ξ = 0, (C12)
eq. (1) and express the hydraulic equation in terms of porosity as
follows: in the invaded region and
 
ct dϕ ∂ k ∂ϕ
= . (C1) f r = 1 at ξ = ∞, (C13)
ϕcϕ dt ∂ x μϕcϕ ∂ x
1526 V. M. Yarushina, D. Bercovici and M. L. Oristaglio

in the reservoir region. At the moving boundary between the regions The leak-off rate is given as
 
f i = f r = f inter , (C14) k ∂ pe  f in−1 k0 c f,i ϕ0 1 d f i 
v= =− √ . (C20)
μi ∂ x x=0 2β μi t dξ ξ =0
  
c f,i d f i c f,r d f r ϕ0 ds √ Substitution of (C16) into (C20) gives the total leak-off
f in−2 = f rn−2 = −2β t = −2βη,
μi dξi μr dξr k0 dt coefficient

(C15) √ 1 k0 c̃i ϕ0
Ct = v t = exp (β p) η̃,
where p = p0 − pfrac > 0 and η > 0 is an unknown constant. β π μi

Similarity variables range between 0 and ξi = c f,i μi η in the in-
√ where
vaded region and between ξr = c f,r μr η and infinity in the reservoir   
region. π μi  
η̃ = 1 − 1 + βη exp c̃i μi η2
When the initial porosity of the reservoir is small (ϕ0
1), (C10) c̃i
can be integrated in a closed form together with the boundary con- 


−1  

Downloaded from https://academic.oup.com/gji/article-abstract/194/3/1514/650635 by guest on 04 July 2020


ditions (C12)–(C14) in the special case n = 1, resulting in the
following porosity distributions within the invaded and reservoir × erf μi c̃i η erf μi c̃i η .
zones:
  

f i = exp (β p) + ( f inter − exp (β p)) erf ci∗ ξi erf μi c̃i η ,


(C16) C2: Non-linear elastic rocks
√ For non-linearly elastic porous rocks in which compaction is de-
at 0 ≤ ξi < c f,i μi η, and
 

 

scribed by a power law, (C5) takes the form


f r = f inter − erf μr c̃r η 1 − erf μr c̃r η − ( f inter − 1)  
∗ −1/q d f d n−1−1/q d f
−2ξ c f = f , (C21)
  

dξ dξ dξ
× erf cr∗ ξr 1 − erf μr c̃r η , (C17)
where

at c f,r μr η < ξr < ∞. To determine the unknown position of the c̃ f 1/q
interface, η, and the unknown porosity at the interface, f inter , we use c∗ = = + 1. (C22)
cf (1 − ϕ0 f ) p0 c f
(C15). The position of the interface is given as an implicit solution
to an algebraic equation: Boundary conditions are:
 f i = ( pfrac / p0 )−q at ξ = 0,
c̃i μr  

π βη 1 − erf μr c̃r η fr = 1 at ξ = ∞,
μi c̃r
and the moving boundary
c̃i √   

= (exp (β p) − 1) − βη πexp c̃i μi η2 erf μi c̃i η . f i = f r = f inter ,


μi
 
(C18) n−2−1/q c f,i d f i c f,r d f r
fi = f rn−2−1/q
μi dξ μr dξ
Porosity at the interface is given by 
  
−1 q ϕ0 ds √ q
π μi   = −2 t = −2η . (C23)
f inter = exp (β p) 1 + βη exp c̃i μi η2 erf μi c̃i η , p0 k0 dt p0
c̃i √
In the invaded zone, ξi = 0 − η c f,i μi , and in the reservoir zone
(C19) √
ξr = η c f,r μr − ∞. Eqs (C21) and (C22) do not have closed form
once η is known. solutions at any values of n and q and are solved numerically.

You might also like