Вы находитесь на странице: 1из 129

A Collection of Recent

Highly Cited
SPE Peer-Reviewed Papers
January 2018
A Collection of R
 ecent Highly Cited 
SPE Peer-Reviewed Papers January 2018
SPE JOURNAL
Shale Gas-in-Place Calculations Part I: New Pore-Scale Considerations
R. J. Ambrose, R. C. Hartman, M. Diaz-Campos, I. Y. Akkutlu, and C. H. Sondergeld. 2012. SPE J 17 (1): 219–229.
SPE-131772-PA. http://doi.org/10.2118/131772-PA.

Recent Advances in Surfactant EOR


G. J. Hirasaki, C. A. Miller, and M. Puerto. 2011. SPE J 16 (4): 889–907. SPE-115386-PA.
http://doi.org/10.2118/115386-PA.

Carbon Dioxide Storage Capacity of Organic-Rich Shales


S. M. Kang, E. Fathi, R. J. Ambrose, I. Y. Akkutlu, and R. F. Sigal. 2011. SPE J 16 (4): 842–855. SPE-134583-PA.
http://doi.org/10.2118/134583-PA.

Simultaneous Multifracture Treatments: Fully Coupled Fluid Flow and Fracture Mechanics
for Horizontal Wells
K. Wu and J. E. Olson. 2015. SPE J 20 (2): 337–346. SPE-167626-PA. http://doi.org/10.2118/167626-PA.

A Generalized Framework Model for the Simulation of Gas Production in Unconventional


Gas Reservoirs
Y.-S Wu, J. Li, D. -Y Ding, C. Wang, and Y. Di. 2014. SPE J 19 (5): 845–857. SPE-163609-PA.
http://doi.org/10.2118/163609-PA.

SPE RESERVOIR EVALUATION & ENGINEERING


Reservoir Modeling in Shale-Gas Reservoirs
C. L. Cipolla, et al. 2010. SPE Res Eval & Eng 13 (4): 638–653. SPE-125530-PA. http://doi.org/10.2118/125530-PA.

Comparison of Fractured-Horizontal-Well Performance in Tight Sand and Shale Reservoirs


E. Ozkan, et al. 2011. SPE Res Eval & Eng 14 (2): 248–259. SPE-121290-PA. http://doi.org/10.2118/121290-PA.

Gas Permeability of Shale


A. Sakhaee-Pour and S. L. Bryant 2012. SPE Res Eval & Eng 15 (4): 401–409. SPE-146944-PA.
http://doi.org/10.2118/146944-PA.

SPE PRODUCTION AND OPERATIONS


What is Stimulated Reservoir Volume?
M. J. Mayerhofer, E. P. Lolon, Cm Rightmire, D. Walser, C. L. Cipolla, and N. R. Warplnskl. 2010. SPE Prod & Oper
25 (1): 89–98. SPE-119890-PA. http://doi.org/10.2118/119890-PA.

Modeling of Hydraulic-Fracture-Network Propagation in a Naturally Fractured Formation


X. Weng, O. Kresse, C. Cohen, R. Wu, and H. Gu. 2011. SPE Prod & Oper 26 (4): 368–380. SPE-140253-PA.
http://doi.org/10.2118/140253-PA.

Learn more about SPE Peer-Reviewed Journals at


http://www.spe.org/publications/
Shale Gas-in-Place Calculations Part I:
New Pore-Scale Considerations
Ray J. Ambrose, Devon Energy and University of Oklahoma; Robert C. Hartman, Weatherford Labs;
and Mery Diaz-Campos, I. Yucel Akkutlu, and Carl H. Sondergeld, University of Oklahoma

Summary When conducting a reservoir study on a natural-gas field, one


Using focused-ion-beam (FIB)/scanning-electron-microscope (SEM) of the primary concerns is the estimation of initial gas in place.
imaging technology, a series of 2D and 3D submicroscale investiga- The estimate is the basis for disclosure of gas reserves, and it is
tions revealed a finely dispersed porous organic (kerogen) material important for reservoir-engineering analysis, such as gas-produc-
embedded within an inorganic matrix. The organic material has tion forecast. Tank-type and multidimensional (simulation-based)
pores and capillaries having characteristic lengths typically less than material-balance calculations are common industry approaches in
100 nm. A significant portion of total gas in place appears to be predicting the gas in place when sufficient field-performance data
associated with interconnected large nanopores within the organic are available. To gain confidence in the estimate, or when sufficient
material. data are not available to initiate the material-balance calculations,
Thermodynamics (phase behavior) of fluids in these pores is the volumetric methods are applied. Using key reservoir param-
quite different; gas residing in a small pore or capillary is rarefied eters (i.e., porosity, water saturation, and formation volume factor)
under the influence of organic pore walls and shows a different associated with well logs, core data, fluid samples, and well tests,
density profile. This raises serious questions related to gas-in-place a volumetric method allows us to predict the gas in place in terms
calculations: Under reservoir conditions, what fraction of the pore of a total-gas pore volume of the reservoir.
volume of the organic material can be considered available as free For disclosure purposes, a deterministic volumetric method, in
gas, and what fraction is taken up by the adsorbed phase? How which a single average value is selected for each parameter in the
accurately is the shale-gas storage capacity estimated using the reserves calculations, is most commonly used in North America.
conventional volumetric methods? And finally, do average densi- Probabilistic methods, on the other hand, are increasingly used
ties exist for the free and the adsorbed phases? worldwide and give the ability to describe the full range of values
We combine the Langmuir adsorption isotherm with the volu- for each parameter in order to somewhat reflect spatial variability
metrics for free gas and formulate a new gas-in-place equation in the parameters and structural intricacies in reservoir architecture.
accounting for the pore space taken up by the sorbed phase. The In this paper, it is shown that any volumetric approach for shale-
method yields a total-gas-in-place prediction. Molecular dynamics gas reserves estimation has added complexity because in shale the
simulations involving methane in small carbon slit-pores of vary- natural gas (mostly methane) exists in different thermodynamic
ing size and temperature predict density profiles across the pores states, namely adsorbed, absorbed (or dissolved), and free gas, and
and show that (a) the adsorbed methane forms a 0.38-nm mono- that an accurate estimation of the gas pore volume in shale reser-
layer phase and (b) the adsorbed-phase density is 1.8–2.5 times voirs should not be considered independently of these states.
larger than that of bulk methane. These findings could be a more A simple model of a physical shale matrix is illustrated in Fig. 1a.
important consideration with larger hydrocarbons and suggest that The model needs to be quantified for gas-in-place analysis, and it
a significant adjustment is necessary in volume calculations, espe- is typically achieved by methods developed specifically for tight
cially for gas shales high in total organic content. Finally, using rocks and other low-permeability formations (Luffel and Guidry
typical values for the parameters, calculations show a 10–25% 1992; Luffel et al. 1993; Mavor and Nelson 1997). However, the
decrease in total gas-storage capacity compared with that using the effective pore volume is not directly determined in these studies;
conventional approach. The role of sorbed gas is more important rather, total porosity, total water volume, and total oil volume
than previously thought. The new methodology is recommended (by weight difference and an assumed oil density of 0.8 g/cc) are
for estimating shale gas in place. determined. Nonetheless, as shown here in Eqs. 1 and 2, the total
and effective gas porosity values are equivalent:
Introduction
Production of natural gas from organic-rich shale makes up an ⎛V ⎞⎛V ⎞ V
increasing portion of the total production in North America. Sg = ⎜ v ⎟ ⎜ g ⎟ = g , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
⎝ Vb ⎠ ⎝ Vv ⎠ Vb
According to the US Department of Energy’s (DOE) Energy
Information Association (EIA), the gross production from shale
plays in the US nearly doubled from 2007 to 2008 from 1,180 Bscf ⎛V ⎞⎛ V ⎞ V
to more than 2,000 Bscf, and has continued to grow (EIA 2009). Sge = ⎜ ve ⎟ ⎜ g ⎟ = g . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
⎝ Vb ⎠ ⎝ Vve ⎠ Vb
The main portion of the shale-gas production is maintained by
the large US shale plays such as the Barnett and Devonian, while
growth is attributable to new production coming on line in the The bulk volume Vb in Eqs. 1 and 2 is determined by mercury
Woodford, Haynesville, Marcellus, Eagle Ford, and Horn River, displacement of competent cores. Grain volume, on the other hand,
to name a few. The natural gas in these reservoirs is stored by two is determined on crushed cores through helium porosimetry. The
mechanisms, free gas and adsorbed gas. The adsorbed gas is stored difference between these two volumes yields the total void volume
mainly on the surface of the organics. Table 1 shows typical total Vv (associated with the total porosity ) available for all in-situ
organic content (TOC) by weight percent (Comer and Hinch 1987; fluids (i.e., mobile hydrocarbons, free and bound water, adsorbed
Jenden et al. 1993; Mancini et al. 2006; Pollastro 2007; Reynolds gas, solution gas, and free gas).
and Munn 2010). For total gas storage, the shale gas-in-place volumes are gener-
ally considered in terms of the following:
• A volumetric component Gf involving hydrocarbons stored in
Copyright © 2012 Society of Petroleum Engineers
the pore space as free gas. The free-gas volume is quantified by
This paper (SPE 131772) was accepted for presentation at the SPE Unconventional modifications of standard reservoir-evaluation methods.
Gas Conference, Pittsburgh, Pennsylvania, USA, 23–25 February 2010, and revised for
publication. Original manuscript received for review 18 January 2010. Revised manuscript • A surface component Ga with the gas physically adsorbed on
received for review 1 December 2010. Paper peer approved 10 December 2010. the large surface area of the micro- and mesopores. The adsorbed-gas

March 2012 SPE Journal 219


In the current industry-standard calculations, Eqs. 6 and 7 are
TABLE 1—TYPICAL TOC OF NORTH AMERICAN SHALE
GAS PLAYS not applied. The solution gas in mobile hydrocarbons and water,
and the adsorbed gas within organic matter, are combined in the
Shale or Play Average or Range of TOC (wt %) adsorption isotherm analysis; therefore, Eq. 3 is reduced to
Barnett 4%
Gst = G f + Ga . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
Marcellus 1–10%
Haynesville 0-8% Note that Gf is equivalent to Eqs. 1 or 2, although, to be consistent
Horn River 3% with the total gas-sorption data Ga, it is now defined in Eq. 4 in
terms of standard cubic feet per ton (Mavor and Nelson 1997).
Woodford 5%
The current volumetric approaches for shale gas are built on
the premise that the two volumes on the right side of Eq. 8, being
associated with the inorganic pores and organic solid, respectively,
amount has generally been quantified from the sorption-isotherm mea- can be estimated independently of one another. Accordingly, the
surements by establishing an equilibrium adsorption isotherm. sorbed gas is associated with the organics, the pore volume of
• A volumetric component Gso involving gas dissolved into which is considered to be negligible, and therefore the volume does
the liquid hydrocarbon. This volume is usually combined with not need to be accounted for during the calculations of free gas;
adsorbed-gas capacity in reservoirs that contain a large fraction of all of the free gas, however, is associated with the inorganic voids
liquid hydrocarbon in the pore space. such as macropores, fissures, fractures, and so on. In this paper,
• A volumetric component Gsw involving gas dissolved into the based on new pore-scale observations, we argue that the total gas-
formation water. The amount of dissolved gas is estimated from storage capacities and the resulting gas-in-place values are being
the bulk-solubility calculations. Although it has traditionally not inadvertently inflated and overestimated by this point of view. The
been considered important, a recent study is available discussing source of the error involves the improper accounting of the pore
significant enhancement in gas solubility in formation liquids when volume occupied by the sorbed-gas phase. There has been some
confined to small pores (Diaz-Campos et al. 2009). awareness of a necessity for the sorbed-phase porosity corrections
Hence, we have Gst as the total gas in place: [see, for example, Cui et al. (2009)], although the issue has not
been addressed properly yet. This article introduces a methodologi-
Gst = G f + Ga + Gso + Gsw , . . . . . . . . . . . . . . . . . . . . . . . . . . (3) cal approach based on new pore-scale considerations.
where the components of storage on the right side are defined as
Sorbed-Phase Correction for the Void Volume
 (1 − Sw − So ) The amount of sorbed gas that is estimated to be in shale is deter-
G f = 32.0368 , . . . . . . . . . . . . . . . . . . . . . . . . (4) mined through an equilibrium adsorption-isotherm experiment. In
b Bg
this experiment, a void volume is first measured, typically using
helium. Void-volume determination is experimentally identical to
p the helium-porosimetry techniques used to determine grain density.
Ga = GsL , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
p + pL Some authors have raised the issue of molecular size of the fluid
used (Bustin et al. 2008; Kang et al. 2010) as a source of error; this
32.0368 So Rso error will not be discussed here. After the void volume has been
Gso = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6) measured, the sorption data are collected. The mass of gas sorbed
5.6146  Bo
into the sample is measured by material balance and a given ther-
modynamic equation of state. During construction of the isotherm,
32.0368 Sw Rsw at each pressure step, the volume of the gas that is adsorbed reduces
Gsw = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
5.6146  Bw the void volume. As a result, the initially determined void volume

Non-Clay Non-Clay
Grain Volume Grain Volume

Dry Clay Dry Clay


Bulk Volume
Bulk Volume

Volume Volume

Bound (Clay) Bound (Clay)


Water Volume Water Volume

Organic Content Organic Content


Sorbed Gas
Effective Void

Volume
Connected Pore Connected Pore
Volume

Total Void
Total Void

Free Gas
Volume
Volume

Volume Containing Volume Containing


Volume
Void

Free Oil, Gas, and Water Free Oil, Gas, and Water

Isolated Pore Volume Isolated Pore Volume

(a) Old Petrophysical Model (b) New Petrophysical Model


Fig. 1—(a) Old petrophysical model and (b) new petrophysical model showing volumetric constituents of a typical gas-shale
matrix. The hatched region describes the interplay between the sorbed phase and total porosity (void volume).

220 March 2012 SPE Journal


50
45

Storage Capacity (scf/ton)


40
35
30
25
20
15
Langmuir Isotherm Data
10
Raw Gibbs Isotherm Data
5
0
0 1000 2000 3000 4000
Pressure (psia)

Fig. 2—Methane isotherms with and without Gibbs correction.

Vv0 must be corrected at the beginning and at the end of the pressure If the changes in void volume are not taken into account, the
step, as described in Eqs. 9 and 10 (Menon 1968): isotherm will be in error and will not be usable in engineering
calculations. An example of isotherm data with and without void-
ˆ
n1 M volume corrections is shown in Fig. 2. The aforementioned void-
Vv1 = Vv 0 − , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9) volume considerations that are necessary to accurately determine
s adsorbed-gas volumes have significant implications on the “live,”
in-situ shale pore volume available for free-gas storage. There-
ˆ
n2 M fore, the effective-porosity/gas-saturation product (Sge) derived
Vv 2 = Vv 0 − . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)
s from a total pore volume that is determined under static labora-
tory conditions does not reflect live-reservoir conditions. The
reservoir total pore volume is not only consumed by water and
Accordingly, over the course of the isotherm analysis, the void
oil but also by the adsorbed gas. It is for this reason we propose
volume is further reduced for each subsequent pressure step. In
that the calculated free-gas pore volume requires a correction for
practice, it is often more convenient to determine a so-called
the adsorbed-gas fraction present under reservoir-temperature and
Gibbs isotherm in terms of number of moles of adsorbed gas, as
-pressure conditions.
in Eq. 11:
Additional evidence to support the need for the correction lies
⎛ pr 1 pr 2 ⎞ ⎛ ps1 ps 2 ⎞ in visual investigation of the complex rock fabric typical of fine-
n2′ = n1′ + Vr ⎜ − + Vv 0 ⎜ − . grained, organic-rich shales. It has recently been reported that
⎝ zr 1 RTr 1 zr 2 RTr 2 ⎠⎟ ⎝ z s1 RTs1 z s 2 RTs 2 ⎠⎟ much of the gas-storage capacity within shale is predominantly
. . . . . . . . . . . . . . . . . . . . . . . (11) associated with the organic fraction of the rock matrix (Loucks
et al. 2009; Wang and Reed 2009; Sondergeld et al. 2010).
The Gibbs isotherm can then be converted to volumes using an Therefore, the free-gas volumes measured by pycnometry using
equation of state and can be adjusted for the void volume using a nonadsorbent gas such as helium must be corrected for the
the Gibbs correction factor f /s: presence of adsorbed gas within the organics. Fig. 3 shows a 2D
FIB/SEM gas-shale image supporting the preceding observations.
Ga' In the image, the matrix is represented by the gray color, with the
Ga = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (12)
 dark- and light-gray regions being the organic (kerogen) and the
1− f inorganic constituents (e.g., clay, silica, and feldspar), respectively,
s
whereas the pores are shown in black. Clearly, most of the pores
are significantly small, typically less than 100 nm, and are almost
exclusively found within the kerogen.
Using advanced imaging technologies, we have created 3D
shale segmentations showing kerogen and pore networks that
contain up to 600 such FIB/SEM images. A similar investigation
has previously been conducted by Tomutsa et al. (2007) for the sub-
micron petrophysical analysis of chalk. A typical kerogen network
of our study is shown in Fig. 4a. The network consists of large
interconnected pockets of kerogen, which make up an estimated
7.7% (volume) of the segment. Fig. 4b shows the superimposed
kerogen (bounded with yellow) and pore (on red) networks of the
same shale segment. As can be seen again from these images,
nearly all of the porosity within the system is associated with the
kerogen network. The image-computed porosity for this segmented
volume is 2.5%. A 3D network overlay showing the comparison
of the kerogen/pore networks and the gray-scale image is shown
in Fig. 3c. Orange denotes kerogen, whereas dark orange denotes
pores. These volumes are produced through gray-scale threshold-
Fig. 3—2D FIB/SEM image showing porosity and kerogen within ing. The setting of thresholds is subjective, and care should be
shale. Black depicts pores; dark gray is kerogen; light gray is taken, although the medium-sized pores are best resolved (Curtis
matrix (clay and silica). et al. 2010).

March 2012 SPE Journal 221


y
z
x
(a) (b) (c)

Fig. 4—3D FIB/SEM segmentations of 300 separate 2D SEM scans. Sample size is 4 ␮m high, 5 ␮m wide, and 2.5 ␮m deep. The
kerogen network is shown in (a); yellow outlines the 3D kerogen network. In (b), yellow outlines the 3D kerogen network, while
red outlines porosity (in this sample, all porosity is found within the kerogen). (c) depicts an overlay of porosity and kerogen 3D
network; orange is kerogen, and dark orange depicts porosity within kerogen. The large orange mass is further in depth than
what is seen on the surface in (a) and (b).

Method for Shale Gas-in-Place Calculations The volume occupied by the sorbed gas must be accounted for after
In this paper, a new petrophysical model is proposed that alters the correction for water saturation. This is because of the porosity
the previous concept of effective porosity discussed earlier. Fig. 1 basis of the water-saturation measurement. Appendix A shows the
shows a comparison between the new model and the old model. derivation of the fundamental equation (Eq. 13). The porosity-
The new model emphasizes two distinct conceptual changes with reduction effects can be shown to vary. In deep reservoirs where
respect to the previous model. First, there exists a dependency of there is not much sorbed-gas content, the correction to the free gas
the connected pore space on the organics (Wang and Reed 2009; in place can be less than 5%. In other reservoirs with low porosity
Loucks et al. 2009; Sondergeld et al. 2010). Second, there is a and a high sorbed-gas content, the reduction could be up to 40%
dependency on the free pore space by the inclusion of a sorbed of the free gas in place or greater. Appendix B shows an example
phase. Fig. 5 shows a simple diagram of the current methodology calculation and the impact on the calculated gas in place.
used to determine gas in place vs. the proposed methodology. For
simplicity, the water and oil volumes are not considered in the Sorbed-Phase Density. In order to calculate the volume occupied
diagram. The simple illustration taken in context with the new by the sorbed phase, the density of adsorbed gas in the organic
information from the FIB/SEM images and segmentations visu- pores (s) must be known a priori. Measurement of the sorbed-
ally shows the errors that result by assuming that the sorbed gas phase density is not a trivial matter, however; the local density for
takes up no volume. methane is expected to vary across the pore and to be different
In order to properly account for the total and free gas in place, from its average bulk density because of the added interactions
the volume occupied by the adsorbed gas must be determined and between methane and the organic pore walls. Furthermore, in gas
subtracted from the free-gas calculation. We therefore propose that shales where the reservoir temperature is significantly greater than
the standard calculations used to calculate free-gas storage capacity the critical temperature of the natural gas, it is difficult to study
(i.e., Eq. 4) be modified as follows: phase transitions and determine if the adsorbate is in the form of
a liquid or vapor.
There have been several suggestions, which have been docu-
32.0368 ⎡  (1 − Sw ) 1.318 × 10 −6 M
ˆ ⎛ p ⎞⎤
Gf = ⎢ − ⎜ GsL ⎥ mented in physical-chemistry literature, to determine the density
Bg ⎣ b s ⎝ p + pL ⎟⎠ ⎦ of the adsorbed phase on solid surfaces. Dubinin (1960) suggested
that the adsorbate density is related to the van der Waals co-volume
. . . . . . . . . . . . . . . . . . . . . . . (13) constant b. Independently, Haydel and Kobayashi (1967) used an
experimental method and found the density values for methane
Old Methodology New Methodology and propane to be nearly equal to the van der Waals co-volume
constant. Later, it was argued that the sorbed-phase density is
Void Space Void Space equivalent to the liquid density (Menon 1968) and to the critical
Measured by Measured by density (Tsai et al. 1985) of the sorbed fluid. Ozawa et al. (1976)
Porosity Porosity considered the adsorbed phase as a superheated liquid with a den-
Measurement Measurement sity dependent upon the thermal expansion of the liquid. Recently
+ Ming et al. (2003) compared all of the previously stated methods
+ Sorbed Mass to a Langmuir-Freundlich adsorption model and found that there
Sorbed Mass Measured by exists a temperature dependence in the sorbed-phase density, but
Measured by Adsorption the value approaches those proposed by Dubinin (1960). These
Adsorption Experiment studies, although fundamentally important to our understanding of
Experiment -
gas adsorption in shale, do not show a clear and accurate path to
Free Gas prediction of the adsorbed-phase density of shale gas and, hence,
Volume Taken to estimation of shale gas in place. All the previous studies were
=Total GIP up by Sorbed
conducted at low pressure and temperature values (at most, 104oF
Gas
=Total GIP and 1,000 psi) or using an inorganic wall (silica gel, as in the
case of Haydel and Kobayashi). In this research, density values
Fig. 5—Comparison of the old and new methodologies in pre- in the first layer within the range 0.28–0.3 g/cc were found using
dicting shale gas-in-place; for simplicity, oil and water volumes molecular modeling and simulation. It is found that the density
are not shown. is sensitive to changes in temperature, pressure, and pore size.

222 March 2012 SPE Journal


Fig. 6—Molecular simulation cell consisting of graphite walls and OPLS-UA methane.

Typically, at 3,043 psi and 176oF, the adsorbed-phase density of MD Simulation of Methane Adsorption in Organic Slit-Pores.
methane is 0.35 g/cc for a pore width of 2.3 nm when the Langmuir For the molecular-level investigation, methane is considered at
single-layer theoretical model is used. some supercritical conditions under thermodynamic equilibrium in
In this study, we used a numerical molecular-modeling approach 3D periodic orthorhombic pore geometry consisting of upper and
to determine the adsorbed-phase density from the first principles lower pore walls made of graphene (carbon) layers (see Fig. 6).
of Newtonian mechanics. Molecular modeling and simulation is a For comparison, we take into account three slit-like pores with a
form of computer simulation that enables us to study thermody- pore width equal to H = 3.9, 2.3, and 1.1 nm. Pore width is an
namic and transport properties of many-particle systems, in which important length scale of the study, and it is defined as the dis-
particles (atoms and molecules that make up the natural gas and tance between the innermost graphene planes. The pores maintain
pore walls) with initially known and instantaneously predicted posi- fixed dimension in the y-coordinate: Ly = 3.93 nm; however, the
tions and momenta are allowed to interact for a period of time, giv- dimension is changed in the x-coordinate such that both pores host
ing a view of the motion of the particles as trajectories in space and approximately the same number of methane molecules (400–450)
time. Numerical integration of Newton’s equations of motion make during the simulations. Hence, the dimensions in this direction are
up the core of the simulation; therefore, special algorithms (e.g., equal to Lx = 4.26 nm and to Lx = 7.67 nm for the large and small
Verlet, Leap Frog, or Beeman algorithms) have been developed and pores, respectively. While the effect of pore pressure is investi-
are commonly used during a molecular simulation study. Readers gated, the number of particles is increased from 400 to 600 within a
who are interested in details of molecular simulation are encour- 2.3-nm pore. Typically, the approximate parallel computation time
aged to consult textbooks by Frenkel and Smit (2002) and Allen was around 1,500–3,000 seconds, and 100 to 144 cores have been
and Tildesley (2002) for deterministic and stochastic treatment used during the simulations.
of the numerical-integration process. Molecular simulations have A united-atom carbon-centered Lennard-Jones potential (based
made enormous strides in recent years and are gradually becoming on the optimized potentials for liquid simulations, OPLS-UA
a commonly used tool in science and engineering (de Pablo and force field) is used as the model of a methane molecule; Table 2
Escobedo 2002). Today, molecular simulations are being widely shows the energy and distance parameters used for fluid/fluid
used to construct virtual experiments in cases where controlled and solid/solid interactions (Nath et al. 1998). The methane/solid
laboratory measurements are difficult, if not impossible, to perform. and methane/methane interactions are of the Lennard-Jones type
There exists an exhaustive literature studying the equilibrium ther- (Frenkel and Smit 2002). A Lorentz-Berthelot mixing rule [ij =
modynamics of fluids using molecular simulation involving phase (ii + jj)/2 and εij = (εiiεjj)1/2] is set up in order to describe solid/
change of bulk fluids (Harris and Yung 1995), characterization of fluid (i.e., pore-wall methane) interactions. Here, ij and εij are
porous materials using gas adsorption (Aukett et al. 1992; Sweat- the Lennard-Jones parameters accounting for interactions between
man and Quirke 2001), and multicomponent gas separation (Ghoufi a molecular site of methane species and a carbon atom of the
et al. 2009). In our case, two sets of molecular-dynamics (MD) organic wall. Lennard-Jones interactions were cut off at 4.1ii for
simulations are used to estimate and analyze the adsorbed-phase the 3.9-nm pore width, and at 3ii for the 1.1- and 2.3-nm pore
density under typical reservoir pressure and temperature conditions: width, respectively. The van der Waals interactions were cut off at
(a) runs involving bulk-phase (i.e., in the absence of pore walls) 3ii for all three systems.
methane-density measurements using a fixed number of methane The MD simulation package, DL-POLY (version 2.20), was
molecules at fixed pressure and temperature and (b) runs involving used to perform the simulations (Todorov and Smith 2008). Meth-
measurements of density of methane at the same temperature but ane bulk-density computations are made considering isobaric and
confined to a pore with organic (graphite) walls. We will consider isothermic conditions with a constant number of atoms, constant
deviations in the methane-density profile along the pore width of pressure, and constant temperature (NPT) at three constant tem-
the second set of runs relative to the uniform density value of the peratures (176, 212, and 266°F) using the Nosé- Hoover thermostat
bulk as an impact of the pore-wall effects. and barostat ensemble (Frenkel and Smit 2002). The relaxation
time used for the Nosé-Hoover thermostat is optimized to 15
pico-seconds (ps). In the case of simulations involving methane in
TABLE 2—LENNARD-JONES POTENTIAL PARAMETERS carbon slit-pores, initially the fluid system was equilibrated at a
FOR METHANE AND CARBON constant temperature, with a canonical ensemble [constant number
of atoms, constant volume, and constant temperature (i.e., NVT)].
Atom (nm) /kB (K) The equilibrium is assumed to be achieved if no drift in time was
Carbon 0.34 28 observed in the time-independent quantity, such as the total energy
of the system. The relaxation time used for the Nosé-Hoover ther-
Methane 0.373 147.9
mostat is optimized to 13, 15, and 20 ps for the pore widths of 1.1,

March 2012 SPE Journal 223


0.4
Pore Width = 3.93nm 0.331
0.30 0.3
0.281
40 4.13E+01 0.2
0.124
0.25 0.1
ρ* CH4 (1000A–3 )

32 0.0
0.20

ρCH4 (g/cc)
0
0.0 3.8 7.6 11.4 15.2 19.
0.163
24
0.15 0.133 0.126 0.124
16
0.10
7.24E+00
8 5.18E+00 0.05
0 0.00
0.0 3.8 7.6 11.4 15.2 19.0 0.0 3.8 7.6 11.4 15.2 19.0
Pore Half Length (Å) Pore Half Length (Å)
Fig. 7—Number-density (left) and discrete-density (right) profile for methane at 176°F (80°C) in a 3.93 nm pore. Density values
are estimated at each 0.2-Å interval for the continuous-density profile. Discrete density corresponds to molecular-layer density
for methane across the pore. The estimated pore pressure at the center of the pores is 3,043 psi. The inset graph in the upper
right corner is the equivalent density using a Langmuir single-layer adsorption model.

2.3, and 3.9 nm, respectively. The Nosé-Hoover algorithm needs to gas density of methane 0.124 g/cc at the center of the pore. The
be used to control the system temperature while the real canonical pore pressure is 3,043 psia, which is a quantity predicted using the
ensemble computations are carried out for the density computa- free-gas molecules and the volume at the center of the pore. The
tions (Frenkel and Smit 2002). The average quantities within observed density profile shows that the assumption of Langmuir
1 nanosecond (ns) intervals after the equilibrium is reached are adsorption theory with monolayer is reasonable to describe the
taken for the analysis. Thermodynamic equilibrium is considered equilibrium adsorption dynamics of methane in the organic pores.
to be reached when the fluctuations in the density profile along the The inset graph on the upper right shows the equivalent Langmuir
time are uniform. Typically, the total run time for a simulation with monolayer plot where the pore half-length has been divided by the
the pore was 1.5 ns and the timestep used was 0.003 ps. adsorbed and free gas. The adsorbed-layer density is estimated as
At the end of each simulation run, number density Number (num- the total of the discrete densities associated with adsorption:
ber of molecules per volume) for methane across the pore space
was computed at every z = 0.2 Å interval (for the continuous s = 0.281 + (0.163–0.124) + (0.133–0.124)
density profile) and at every Lz = 3.8 Å (for the discrete density + (0.126–0.124) = 0.33g/cm3.
profile) volume segment in the z-direction. The number density
for each volume segment is estimated by summing the number of Pore-Size Effects on Methane Adsorption. In essence, super-
molecules counted in each interval within the volume segment and critical methane in small organic pores is structured as a result
dividing the total by the number of intervals. Then, the number of pore-wall effects, showing layers of graded density across the
density is converted to the local mass density of methane using its pore. Depending on the pore size, a bulk-fluid region may exist at
molecular weight MCH4 and Avogadro’s number as follows: the central portion of the pore, where the influence of molecular
interactions with the pore walls is very small. In pores with sizes up
 Number M CH4 to 50 nm, a combination of molecule/molecule and molecule/wall
CH4 = .
interactions dictates thermodynamic states of the gas and its mass
6.02252 × 10 23
transport in the pore (Krishna 2009). On the other hand, within a
As explained earlier, it is expected that the local mass density small pore (see Fig. 8), methane molecules are always under the
for methane across the pore will be different from its mean bulk influence of the force field exerted by the walls; consequently, no
density, owing to changing levels of interactions between the bulk-fluid region can be observed in the pore, and no pore pressure
methane/methane and methane/carbon bodies. The purpose of our can be measured. Therefore, behavior of the adsorbed molecules
numerical investigation was to predict a precise density profile should be considered rather than the motion of free-gas molecules.
across the pore as an indication of the presence of these interac- Furthermore, the adsorbed-phase density at the first layer is signifi-
tions and to determine an average adsorbed-gas-density value, a cantly large. The smaller the pore, the larger the adsorbed-phase
macroscopic quantity necessary for the gas-in-place calculations density is. In the limit, Langmuir theory is not suitable for the
using Eq. 13. description of physical interactions between the gas and solid.
Fig. 7 shows the density profile for methane confined to the Effect of Temperature on Methane Adsorption. The effect
3.92-nm pore at 176°F. It is clear that the predicted methane density of temperature on methane density in the large pore is shown in
is not uniform across the pores; its value is significantly greater Fig. 9. The estimated average adsorbed-methane density in the
near the wall, where adsorption takes place, and it decreases with first layer is around 0.3 g/cm3, although it decreases with tem-
damped oscillations as the distance from the pore wall increases. perature (for example, to 0.259 g/cm3 at 266°F). These values
The oscillations are caused by the presence of adsorption in show variations caused by changing levels of kinetic energy at
the pores and involve structured distribution of molecules (i.e., the microscopic scale.
molecular layers). The number of molecules is largest in the first Table 3 shows that the predicted free-gas density values at the
layer near the wall, indicating physical adsorption. The molecules center of the pore correspond to the bulk methane density values
in the second layer are still under the influence of pore walls, predicted independently using the National Institute of Standards
although intermolecular interactions among the methane molecules and Technology’s (NIST) thermophysical properties of hydro-
begin to dominate, not allowing locally high methane densities. In carbon mixtures database, SUPERTRAPP. The comparisons also
this layer, the density of methane is slightly larger than the bulk show that our numerical simulations are accurate.

224 March 2012 SPE Journal


0.4 0.350
Pore Width = 2.31nm 0.40
0.3

50 0.35 0.2
0.124
0.1
42.80 0.30 0.292
43.01 0.0
40
0.25 0.0 3.8 7.6

ρCH4 (g/cc)
ρ* CH4 (1000Å–3 )

30 0.20 0.171
0.15 0.135
20
0.10
10 7.65 7.72
0.05
4-5 (bulk)
0 0.00
0.0 3.8 7.6 11.4 15.2 19.0 22.8 0.0 3.8 7.6 11.4
Pore Length (Å) Pore Half Length (Å)
Pore Width = 1.14nm

50 4.57E+01 0.35
0.307
4.69E+01 0.30
40
ρ* CH4 (1000Å–3 )

ρCH4 (g/cc) 0.25 0.234


30 0.20

20 0.15
1.57E+01
0.10
10
0.05
4-5(bulk)
0 0.00
0.0 3.8 7.6 11.4 0.0 3.8
Pore Length (Å) Pore Half Length (Å)
Fig. 8—Number-density (left) and discrete-density (right) profiles for methane at 176°F (80°C) in pore widths of 2.31 nm (upper)
and 1.41 nm (lower). Density values are estimated at each 0.2-Å interval for the continuous-density profile. Discrete density cor-
responds to molecular-layer density for methane across the pore. Estimated pore pressure at the center of the pores is 3,043
psi. The inset graph in the upper right corner is the equivalent density using a Langmuir single-layer adsorption model.

Conclusions adjustments. This change in gas in place persists even when the
In this paper, we addressed the following issues in regard to the sorbed density values reported by other researchers, such as the
volume available for free gas in organic shales: liquid-methane density of 0.4223 g/cc, corresponding to 1 atm
• The sorbed phase follows Langmuir theory for most of the pores and –161°C (Mavor et al. 2004), and of 0.374 g/cc (Haydel and
and takes up a one-molecule-thick portion of a pore, although Kobayashi et al. 1967). These are clearly shown in Table 4.
there is a damped oscillation density profile with an increased In conclusion, a robust method that matches the local phys-
density in the second layer. For a 100-nm pore, the volume is ics is presented to determine an estimate of the gas in place in
fairly insignificant; however, for pores on the order of 1 nm, it organic-rich gas shale. Future work includes re-evaluation of the
is quite large. concepts and approaches that are presented and discussed in the
• The current industry standard disregards the volume consumed presence of multicomponent gases with varying pore sizes. It is
by the sorbed phase, thus inadvertently overestimating the pore expected that there will be added uncertainties in the gas-in-place
volume available for free-gas storage. We showed that with Eq. estimation because of (1) multicomponent nature of gas and its
13, a more correct volume is calculated. Through MD simulation adsorbed phase and (2) volume fraction of organic nanopores that
and Langmuir theory, we showed that the density for adsorbed are currently inaccessible for visual observation and measurement.
methane typically equals 0.34 g/cm3. This value is based on The former requires computer-simulation studies on the adsorbed-
numerical work using simple flat organic pore-wall surfaces. In phase thickness and density using binary, ternary, and quaternary
small pores, the adsorbed-layer thickness will be affected by the mixtures of gases with components such as ethane, propane, and
roughness and curvature of the solid surface, and the density will carbon dioxide. The latter uncertainty, on the other hand, can be
increase because of increased surface area of the wall. Further reduced by measuring the absolute area of the organic surfaces
analysis using more-complex pore-wall structures is necessary. using isotherm data. In addition, a new study on transport proper-
• In our calculation of the gas in place, we predict a dramatic ties of natural gases in nanoporous organic materials is necessary
change in the estimated gas-in-place values because of volume in light of the new pore-scale findings.

March 2012 SPE Journal 225


40 0.30
176°F 176°F
0.25 212°F
32 212°F
266°F
ρ* CH4 (1000Å–3 )

266°F
0.20

ρCH4 (g/cc)
24
0.15
16
0.10
8
0.05

0 0.00
0.0 3.8 7.6 11.4 15.2 19.0 0.0 3.8 7.6 11.4 15.2 19.0
Pore Half Length (Å) Pore Half Length (Å)

Fig. 9—Number-density (left) and discrete-density (right) profiles across half-length of a 3.9-nm-width slit-pore as a function of
temperature.

TABLE 3—EFFECT OF PRESSURE ON METHANE DENSITY


Methane Bulk Density Methane Bulk Density*
3 3
Temperature (°F) Pressure (psi) (g/cm ) (g/cm )

176 2206 0.090 0.089


176 3043 0.124 0.122
176 3676 0.147 0.144
176 4141 0.160 0.159
176 4404 0.168 0.167
176 4586 0.176 0.172
176 4800 0.178 0.175
176 6272 0.214 0.211
176 7300 0.231 0.229
176 7550 0.235 0.233
176 8707 0.253 0.250
NOTE: The estimated density values are compared with the bulk density values obtained using
SUPERTRAPP® under the same pressure and temperature conditions.
* From NIST-SUPERTRAPP®

Nomenclature Ga = adsorbed-gas storage capacity, scf/ton


Bg = gas formation volume factor, reservoir volume/surface Gf = free-gas storage capacity, scf/ton
volume GsL = Langmuir storage capacity, scf/ton
Bo = oil formation volume factor, reservoir volume/surface Gso = dissolved-gas-in-oil storage capacity, scf/ton
volume Gst = total gas storage capacity, scf/ton
Bw = water formation volume factor, reservoir volume/surface Gsw = dissolved-gas-in-water storage capacity, scf/ton
volume H = pore size, Å

TABLE 4—EFFECT OF SORBED PHASE DENSITY ON FREE AND TOTAL GIP CALCULATIONS
Adsorbed-Phase Density Free Storage Capacity Total Storage Capacity
3
(g/cm ) (SCF/ton) (SCF/ton)

Reference s Shale A Shale B Shale A Shale B

Molecular dynamics 0.34 91.9 75.8 130.7 158.6


Haydel and Kobayashi (1967) 0.37 72.9 95.1 155.7 133.9
Mavor et al. (2004)
(at 1 atm, –161.2°C) 0.42 93.2 79.7 132.1 162.5

226 March 2012 SPE Journal


kB = Boltzmann constant, 1.3806503×10–23 kJ/K–1 Bustin, R.M., Bustin, A.M., Cui, X., Ross, D.J.K., and Murthy Pathi, V.S.
Lx = length of computational cell in x-direction, Å 2008. Impact of Shale Properties on Pore Structure and Storage Charac-
Ly = length of computational cell in y-direction, Å teristics. Paper SPE 119892 presented at the SPE Shale Gas Production
Conference, Fort Worth, Texas, USA, 16–18 November. http://dx.doi.
M̂ = apparent natural-gas molecular weight, lbm/lbmole
org/10.2118/119892-MS.
MCH4 = molecular weight of methane, g/gmole Comer, J.B. and Hinch, H.H. 1987. Recognizing and quantifying expul-
n = number of moles sion of oil from the Woodford Formation and age-equivalent rocks in
n’1 = Gibbs-isotherm number of moles at the start of pressure Oklahoma and Arkansas. AAPG Bulletin 71 (7): 844–858.
step, lb-moles Cui, X., Bustin, A.M.M., and Bustin, R.M. 2009. Measurements of
n’2 = Gibbs-isotherm number of moles at the end of pressure gas permeability and diffusivity of tight reservoir rocks: different
step, lb-moles approaches and their applications. Geofluids 9 (3): 208–223. http://
p = pressure, psia dx.doi.org/10.1111/j.1468-8123.2009.00244.x.
pL = Langmuir pressure, psia Curtis, M.E., Ambrose, R.J., Sondergeld, C.S., and Rai, C.S. 2010.
pr1 = pressure of reference cell at start, psia Structural Characterization of Gas Shales on the Micro- and Nano-
pr2 = pressure of reference cell at end, psia Scales. Paper SPE 137693 presented at the Canadian Unconventional
ps1 = pressure of sample cell at start, psia Resources and International Petroleum Conference, Calgary, 19–21
October. http://dx.doi.org/10.2118/137693-MS.
ps2 = pressure of sample cell at end, psia
de Pablo, J.J. and Escobedo, F.A. 2002. Molecular Simulations in Chemical
R = universal gas constant, psia-ft3/mol-K Engineering: Present and future. AIChE Journal 48 (12): 2716–2721.
Rso = solution-gas/oil ratio, scf/STB http://dx.doi.org/10.1002/aic.690481202.
Rsw = solution-gas/water ratio, scf/STB Diaz-Campos, M., Akkutlu, I.Y., and Sigal, R.F. 2009. A Molecular
Sg = gas saturation, dimensionless Dynamics Study on Natural Gas Solubility Enhancement in Water Con-
Sge = effective gas saturation, dimensionless fined to Small Pores. Paper SPE 124491 presented at the SPE Annual
So = oil saturation, dimensionless Technical Conference and Exhibition, New Orleans, 4–7 October.
Sw = water saturation, dimensionless http://dx.doi.org/10.2118/124491-MS.
T = reservoir temperature, °F Dubinin, M.M. 1960. The Potential Theory of Adsorption of Gases and
= reference-cell temperature at start, °R Vapors for Adsorbents with Energetically Nonuniform Surfaces. Chem-
Tr1
ical Review 60 (2): 235–241. http://dx.doi.org/10.1021/cr60204a006.
Tr2 = reference-cell temperature at end, °R
Energy Information Administration (EIA). 2009. Annual Energy Out-
Ts1 = sample-cell temperature at start, °R look 2009, With Projections to 2030. Annual Report No. DOE/EIA-
Ts2 = sample-cell temperature at end, °R 0383(2009), EIA/US DOE, Washington, DC (March 2009).
Vb = bulk volume, ft3 Frenkel, D. and Smit B. 2002. Understanding Molecular Simulation: From
Vg = gas volume, ft3 Algorithms to Applications, second edition, Vol. 1. San Diego, Califor-
Vr = reference volume, ft3 nia: Computational Science Series, Academic Press.
Vv = void volume, ft3 Ghoufi, A., Gaberova, L., Rouquerol, J., Vincent, D., Llewellyn, P.L., and
Vve = effective void volume, ft3 Maurin, G. 2009. Adsorption of CO2, CH4 and their binary mixture in
Vv0 = initial void volume, ft3 Faujasite NaY: A combination of molecular simulations with gravime-
Vv1 = void volume Step 1, ft3 try–manometry and microcalorimetry measurements. Microporous and
Mesoporous Materials 119 (1–3): 117–128. http://dx.doi.org/10.1016/
Vv2 = void volume Step 2, ft3
j.micromeso.2008.10.014.
z = compressibility factor Harris, J.G. and Yung, K.H. 1995. Carbon Dioxide’s Liquid-Vapor Coex-
zr = compressibility factor of reference cell istence Curve And Critical Properties as Predicted by a Simple Molec-
z = interval across the pore space used for number density ular Model. J. Phys. Chem. 99 (31): 12021–12024. http://dx.doi.
calculations, Å org/10.1021/j100031a034.
ε = depth of the potential well, kJ Haydel, J.J. and Kobayashi, R. 1967. Adsorption Equilibria in the Methane-
b = bulk-rock density, g/cm3 Propane-Silica Gel System at High Pressures. Ind. Eng. Chem. Funda-
CH4 = mass density of methane in pore, g/cm3 men. 6 (4): 564–554. http://dx.doi.org/10.1021/i160024a010.
f = free-gas-phase density, g/cm3 Hünenberger, P.H. 2005. Thermostat Algorithms for Molecular Dynamics
Number = number density of methane, number of molecules/Å3 Simulations. Adv. Polym. Sci. 173: 105–149. http://dx.doi.org/10.1007/
s = sorbed-phase density, g/cm3 b99427.
Jenden, P.D., Drazan, D.J., and Kaplan, I.R. 1993. Mixing of Thermogenic
 = distance at which the intermolecular potential is zero, nm
Natural Gases in Northern Appalachian Basin. AAPG Bulletin 77 (6):
 = total porosity fraction, dimensionless 980–998.
a = sorbed-phase porosity fraction, dimensionless Kang, S., Fathi, E., Ambrose, R.J., Akkutlu, I.Y., and Sigal, R.F. 2010.
e = effective porosity fraction, dimensionless Carbon Dioxide Storage Capacity of Organic-rich Shales. Paper
SPE 134583 presented at the SPE Annual Technical Conference
Acknowledgments and Exhibition, Florence, Italy, 19–22 September. http://dx.doi.
We recognize the work and contributions of Mark Curtis and Gary org/10.2118/134583-MS.
Stowe in collecting many of the images in this paper. We thank the Krishna, R. 2009. Describing the Diffusion of Guest Molecules Inside
staff at the University of Oklahoma’s OSCER supercomputing cen- Porous Structures. J. Phys. Chem. C 113 (46): 19756–19781. http://
ter for the facilities to perform the MD simulations. We also thank dx.doi.org/10.1021/jp906879d.
Devon Energy—in particular, Jerry Youngblood, Bret Jameson, Loucks, R.G., Reed, R.M., Ruppel, S.C., and Jarvie, D.M. 2009. Morphol-
Jeff Hall, and Bill Coffey—for providing access to Barnett core ogy, Genesis, and Distribution of Nanometer-Scale Pores in Siliceous
and supporting much of this work. Mudstones of the Mississippian Barnett Shale. Journal of Sedimentary
Research 79 (12): 848–861. http://dx.doi.org/10.2110/jsr.2009.092.
References Luffel, D.L. and Guidry, F.K. 1992. New Core Analysis Methods for
Allen, M.P., and Tildesley, D.J. 2002. Computer Simulation of Liquids, Measuring Rock Properties of Devonian Shale. J. Pet. Tech. 44 (11):
reprint. London: Oxford University Press. 1184–1190. SPE-20571-PA. http://dx.doi.org/10.2118/20571-PA.
Aukett, P.N., Quirke, N., Riddiford, S., and Tennison, S.R. 1992. Meth- Luffel, D.L., Hopkins, C.W., and Schettler, P.D. Jr. 1993. Matrix Permeabil-
ane adsorption on microporous carbons—A comparison of experi- ity Measurement of Gas Productive Shales. Paper SPE 26633 presented
ment, theory, and simulation. Carbon 30 (6): 913–924. http://dx.doi. at the SPE Annual Technical Conference and Exhibition, Houston, 3–6
org/10.1016/0008-6223(92)90015-O. October. http://dx.doi.org/10.2118/26633-MS.

March 2012 SPE Journal 227


Mancini, E. A., Goddard D. A., Barnaby R., and Aharon P. 2006. Basin V RT
Analysis and Petroleum System Characterization and Modeling, Inte- =
n p
rior Salt Basins, Central and Eastern Gulf of Mexico. Phase 1 Final
Report, Project No. 422, Contract No. DE-FC26-03NT15395, US ft 3 psi
10.73159 519.67˚R
DOE/NETL, Alabama University, Tuscaloosa, Alabama. V ˚R lb-mol . . . . . . . . . . . . . . . . . . (A-2)
=
Mavor, M.J. and Nelson, C.R. 1997. Coalbed Reservoir Gas-In-Place n ft 3
14.696 psia = 379.48
Analysis. Report GRI-97/0263, Gas Research Institute, Chicago, lb-mol
Illinois.
Mavor, M.J., Hartman, C., and Pratt, T.J. 2004. Uncertainty in Sorption With density in g/cm3 and the desired units in scf/ton, we can use
Isotherm Measurements. Paper No. 411 presented at the International the preceding value to calculate a conversion constant.
Coalbed Methane Symposium, Tuscaloosa, Alabama, 14–16 May.
Menon, P.G. 1968. Adsorption at high pressures. Chem. Rev. 68 (3): 1 1 ton ton mol . . . . . . . . . (A-3)
⋅ = 1.318 × 10 −6
277–294. http://dx.doi.org/10.1021/cr60253a002. ft 3 2000 lb ft 3
Ming, L., Anzhong, G., Xuesheng, L., and Rongshun, W. 2003. Deter- 379.48
lb-mol
mination of the adsorbate density from supercritical gas adsorption
equilibrium data. Carbon 41 (3): 585–588. http://dx.doi.org/10.1016/ Using the conversion constant, the density of the adsorbed phase,
S0008-6223(02)00356-1. the bulk density of the rock, and the molecular weight of the
Nath, S.K, Escobedo, F.A., and de Pablo, J.J. 1998. On the simulation of adsorbed phase, we can calculate the fractional volume occupied
vapor-liquid equilibria for alkanes. J. Phys. Chem. 108 (23): 9905– by the sorbed phase:
9911. http://dx.doi.org/10.1063/1.476429.
Ozawa, S., Kusumi, S., and Ogino, Y. 1976. Physical adsorption of gases at
b ⎛ p ⎞
high pressure. IV. An improvement of the Dubinin—Astakhov adsorp- a = 1.318 × 10 −6 Mˆ G . . . . . . . . . . . . . . . . . (A-4)
tion equation. Journal of Colloid Interface Science 56 (1): 83–91. s ⎜⎝ sL p + pL ⎟⎠
http://dx.doi.org/10.1016/0021-9797(76)90149-1.
Pollastro, R.M. 2007. Total petroleum system assessment of undiscovered Assuming the oil saturation is negligible, and taking the frac-
resources in the giant Barnett Shale continuous (unconventional) tional sorbed-phase volume away from the free-gas volume, Eq.
gas accumulation, Fort Worth Basin, Texas. AAPG Bulletin 91 (4): 4 becomes
551–578. http://dx.doi.org/10.1306/06200606007.
Reynolds, M.M. and Munn, D.L. 2010. Development Update for an
Emerging Shale Gas Giant Field—Horn River Basin, British Columbia,  (1 − Sw ) − a
G f = 32.0368 . . . . . . . . . . . . . . . . . . . . . . . (A-5)
Canada. Paper SPE 130103 presented at the SPE Unconventional Gas b Bg
Conference, Pittsburg, Pennsylvania, USA, 23–25 February. http://
dx.doi.org/10.2118/130103-MS. Substituting the expression for a into this equation and simplify-
Sondergeld, C.H., Ambrose, R.J., Rai, C.S., and Moncrieff, J. 2010. Micro- ing yields Eq. 13:
Structural Studies of Gas Shales. Paper SPE 131771 presented at the
SPE Unconventional Gas Conference, Pittsburg, Pennsylvania, USA, 32.0368 ⎡  (1 − Sw ) 1.318 × 10 −6 M
ˆ ⎛ p ⎞⎤
23–25 February 2010. http://dx.doi.org/10.2118/131771-MS. Gf = ⎢ − ⎜⎝ GsL p + p ⎟⎠ ⎥
Bg ⎣ b s L ⎦
Sweatman, M.B. and Quirke, N. 2001. Characterization of Porous Materi-
als by Gas Adsorption: Comparison of Nitrogen at 77 K and Carbon
Dioxide at 298 K for Activated Carbon. Langmuir 17 (16): 5011–5020. . . . . . . . . . . . . . . . . . . . . . . (A-6)
http://dx.doi.org/10.1021/la010308j.
Todorov, I.T. and Smith, W. 2008. The DLPOLY_3 User Manual. Cheshire, Appendix B—Example Calculation
UK: Science & Technology Facilities Council (STFC). In order to quantify the pore-scale effects using the new methodol-
Tomutsa, L., Silin, D., and Radmilovic V. 2007. Analysis of Chalk Petro- ogy, we compare the results of the old method vs. the new method
physical Properties by Means of Submicron-scale Pore Imaging and on two shales. The first shale has a low sorbed-gas volume, while
Modeling. SPE Res Eval & Eng 10 (3): 285–293. SPE-99558-PA. the second has a relatively high sorbed-gas volume. A value of 0.34
http://dx.doi.org/10.2118/99558-PA. g/cm3 is used for sorbed-methane density based on the discussion
Tsai, M.C., Chen, W.N., Cen, P.L., Yang, R.T., Kornosky, R.M., Holcombe, in the “MD Simulation of Methane Adsorption in Organic Slit-
N.T., and Strakey, J.P. 1985. Adsorption of gas mixture on acti- Pores” subsection. The rest of the parameters for the two shales
vated carbon. Carbon 23 (2): 167–73. http://dx.doi.org/10.1016/0008- are shown in Table B-1.
6223(85)90008-9. Using the old calculation method for gas in place, the adsorbed,
Wang, F.P. and Reed, R.M. 2009. Pore Networks and Fluid Flow in Gas free, and total gas storage capacities for Shale A are, respectively,
Shales. Paper SPE 124253 presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, 4–7 October. http://dx.doi.
 (1 − Sw )
org/10.2118/124253-MS. G f = 32.0368
Yee, D., Seidle, J.P., and Hanson, W.B. 1993. Gas sorption on coal and b Bg
measurement of gas content. In Hydrocarbons from Coal, ed. B.E. Law 0.06(1 − 0.35)
and D.D. Rice, No. 38, Chap. 9, 203–218. Tulsa, Oklahoma: AAPG = 32.0368 = 108.6 scf/ton . . . . . . . . . . . (B-1)
2.5 ⋅ 0.0046
Studies in Geology Series, AAPG.

Appendix A—Derivation of Eq. 13


p 4000
Beginning with Eq. 5, the value Ga needs to be converted into a Ga = GsL = 50 = 38.8 scf/ton . . . . . . . (B-2)
volume, and a simple unit conversion can be performed. The typi- p + pL 4000 + 1150
cal unit of the equation below is scf/ton.

Gst = G f + Ga = 108.6 scf/ton + 38.8 scf/ton = 147.7 scf/ton


p
Ga = GsL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-1)
p + pL
. . . . . . . . . . . . . . . . . . . . . . (B-3)

Given that it is in scf, we can convert scf into a mass with the Using the new calculation method for gas in place, the free, sorbed,
ideal-gas law at standard temperature and pressure: and total storage capacities for Shale A are, respectively,

228 March 2012 SPE Journal


TABLE B-1—SHALE PROPERTIES FOR GAS-IN-PLACE EXAMPLE CALCULATION

Shale A (Low Sorption Capacity) Shale B (High Sorption Capacity)

φ 0.06 0.06
Sw 0.35 0.35
So 0.0 0.0
Bg 0.0046 0.0046
M̂ 16 lb/lb-mol 16 lb/lb-mol
GsL 50 scf/ton 120 scf/ton
p 4000 psia 4000 psia
T
o o
180 F 180 F
pL 1150 psia 1800 psia
3 3
b 2.5 g/cm 2.5 g/cm
3 3
s
0.34 g/cm 0.34 g/cm

32.0368 ⎡  (1 − Sw ) 1.318 × 10 −6 M
ˆ ⎛ p ⎞⎤ Ga = 120
4000
= 82.8 scf/ton . . . . . . . . . . . . . . . . (B-11)
Gf = ⎢ − ⎜ GsL ⎥
Bg ⎣ b s ⎝ p + pL ⎟⎠ ⎦ 4000 + 1800

⎡ 0.06(1 − 0.335) ⎤ Gst = 71.8 scf/ton + 82.8 scf/ton


32.0368 ⎢ 2.5 ⎥
= ⎢ −6 ⎥ = 155.7 scff/ton. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-12)
0.0046 ⎢ 1.318 × 10 ⋅ 16 ⎛ 4000 ⎞⎥
− ⎜ 50 ⎟
⎢⎣ 0.34 ⎝ 4000 + 1150 ⎠ ⎥⎦ This represents a decrease of 32.8 and 18.6% of the free and total
= 91.9 scf/ton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-4) gas-storage capacities, respectively.

Raymond J. Ambrose is the Director of Reservoir Engineering


4000 for Reliance Holding USA, Inc. His current research interests
Ga = 50 = 38.8 scf/ton . . . . . . . . . . . . . . . . . . (B-5)
4000 + 1150 include analytical solutions for shale productivity, scanning
electron microscopy and pore structure characterization,
and storage mechanisms for shale. He holds a BS degree in
Gst = 91.9 scf/ton + 38.8 scf/ton = 130.7 scf/ton
n . . . . . . . . (B-6) chemical engineering and an MSc degree in petroleum engi-
neering, both from the University of Southern California, and a
This represents a decrease of 15.3 and 11.5% of the free and total PhD degree in petroleum engineering from the University of
Oklahoma. Ambrose currently serves SPE as a Technical Editor
gas-storage capacities, respectively. for the Journal of Canadian Petroleum Technology.
Using the old calculation method for gas in place, the adsorbed,
free, and total gas storage capacities for Shale B are, respectively, Chad Hartman is Chief Technical Advisor for Weatherford
Laboratories, based in Golden, Colorado. Hartman is acknowl-
edged as an industry expert in unconventional reservoirs and
 (1 − Sw ) integrating formation evaluation data. His credentials include
G f = 32.0368
b Bg design and development of Weatherford Laboratories’ state-
of-the-art Adsorption Isotherm Laboratory, field and laboratory
0.06(1 − 0.35)
= 32.0368 = 108.6 scf/ton . . . . . . . . . . . (B-7) protocols, and data QC/QA methods. He currently supports
2.5 ⋅ 0.0046 Weatherford Laboratories’ global business development, prod-
uct research and development, data interpretation, and client
consultation services. He holds a BS degree in chemistry from
p Fort Lewis College.
Ga = GsL
p + pL
Mery Diaz-Campos is with Schlumberger Information Solutions
4000 in Houston. She holds an MSc degree in petroleum engineering
= 120 = 82.8 scf/ton . . . . . . . . . . . . . . . . . (B-8) from the University of Oklahoma and a BSc degree in chemis-
4000 + 1800
try from National University in Lima, Peru.
I. Yucel Akkutlu is an associate professor of petroleum engi-
Gst = G f + Ga neering at the University of Oklahoma. His current academic
= 108.6 scf/ton + 82.8 scf/ton. . . . . . . . . . . . . . . . . . . (B-9) research is on thermodynamics and transport of fluids in nano-
porous materials, and on scaling-up and homogenization of
= 191.4 scf/ton coupled transport and reaction processes in low-permeability
geological formations exhibiting multiscale pore structures. He
Using the new calculation method for gas in place, the free, sorbed, holds MSc and PhD degrees in petroleum engineering from
and total storage capacities for Shale B are, respectively, the University of Southern California. Akkutlu currently serves
SPE as an Associate Editor of SPE Journal.
Carl H. Sondergeld is currently a professor and the Curtis
⎡ 0.06(1 − 0.35) ⎤ Mewbourne Chair at the Mewbourne School of Petroleum and
32.0368 ⎢ 2.5 ⎥
⎢ ⎥ Geological Engineering at the University of Oklahoma. He spent
Gf = −6
0.0046 ⎢ 1.318 × 10 ⋅ 16 ⎛ 4000 ⎞⎥ 19 years at the Tulsa Research Center of Amoco Production
− ⎜⎝ 120 ⎟ Company and holds 14 US patents. He holds BA and MA
⎢⎣ 0.34 4000 + 1800 ⎠ ⎥⎦ degrees in geology from Queens College, City University of New
= 72.9 scf/ton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-10)) York, and a PhD degree in geophysics from Cornell University.

March 2012 SPE Journal 229


Recent Advances in Surfactant EOR
George J. Hirasaki, SPE, Clarence A. Miller, SPE, and Maura Puerto, Rice University

Summary revival of interest, it seems appropriate to review understanding of,


In this paper, recent advances in surfactant enhanced oil recovery and prospects for, surfactant EOR.
(EOR) are reviewed. The addition of alkali to surfactant flooding Adding surfactant to injected water to reduce oil/water IFT
in the 1980s reduced the amount of surfactant required, and the and/or alter wettability and thereby increase recovery is not a new
process became known as alkaline/surfactant/polymer flooding idea [see, for instance, Uren and Fahmy (1927)]. Indeed, a few
(ASP). It was recently found that the adsorption of anionic surfac- early field trials where small amounts of surfactant were injected
tants on calcite and dolomite can also be significantly reduced with did produce small increases in oil recovery. The increases were
sodium carbonate as the alkali, thus making the process applicable probably caused mainly by wettability changes, although the data
for carbonate formations. The same chemicals are also capable were inconclusive for assessing mechanisms. The results were not
of altering the wettability of carbonate formations from strongly sufficiently promising to stimulate use of surfactants on a larger
oil-wet to preferentially water-wet. This wettability alteration in scale. A related long-held concept for improving recovery is to
combination with ultralow interfacial tension (IFT) makes it pos- generate surfactant in situ by injecting an alkaline solution (Atkin-
sible to displace oil from preferentially oil-wet carbonate matrix son 1927), which is less expensive than synthetic surfactants and
to fractures by oil/water gravity drainage. converts naphthenic acids in the crude oil to soaps. Early results
The alkaline/surfactant process consists of injecting alkali and were not encouraging, and the relative importance of likely process
synthetic surfactant. The alkali generates soap in situ by reaction mechanisms was not understood (Johnson 1976). Other references
between the alkali and naphthenic acids in the crude oil. It was to early work on surfactants are given by Hill et al. (1973).
recently recognized that the local ratio of soap/surfactant deter- Two different approaches stimulated significant advances in
mines the local optimal salinity for minimum IFT. Recognition of surfactant EOR processes in the 1960s. The surfactants were made
this dependence makes it possible to design a strategy to maximize either by direct sulfonation of aromatic groups in refinery streams
oil recovery with the least amount of surfactant and to inject poly- or crude oils, or by organic synthesis of alkyl/aryl sulfonates,
mer with the surfactant without phase separation. An additional which allowed for the surfactant to be tailored to the reservoir of
benefit of the presence of the soap component is that it generates interest. The advantages of these surfactants are their low cost, their
an oil-rich colloidal dispersion that produces ultralow IFT over a wide range of properties, and the availability of raw materials in
much wider range of salinity than in its absence. somewhat large quantities.
It was once thought that a cosolvent such as alcohol was Miscible flooding was an active area of research, but the sol-
necessary to make a microemulsion without gel-like phases or vents being considered, such as enriched gas and LPG, exhibited
a polymer-rich phase separating from the surfactant solution. poor reservoir sweep because the adverse mobility ratio promoted
An example of an alternative to the use of alcohol is to blend viscous fingering and the low solvent density led to gravity over-
two dissimilar surfactants: a branched alkoxylated sulfate and a ride. Seeking a solvent miscible with oil but having a higher vis-
double-tailed, internal olefin sulfonate. The single-phase region cosity and density, Gogarty and coworkers at Marathon proposed
with NaCl or CaCl2 is greater for the blend than for either sur- using a slug of an oil-continuous microemulsion made of hydro-
factant alone. It is also possible to incorporate polymer into such carbon, a petroleum sulfonate surfactant, an alcohol, and water
aqueous surfactant solutions without phase separation under some or brine [see review by Gogarty (1977)]. Holm and coworkers at
conditions. The injected surfactant solution has underoptimum Union Oil advocated a similar process using a “soluble oil,” which
phase behavior with the crude oil. It becomes optimum only as was also an oil-continuous microemulsion made mainly of crude
it mixes with the in-situ-generated soap, which is generally more oil, some mineral oil, petroleum sulfonate, a cosolvent such as eth-
hydrophobic than the injected surfactant. However, some crude ylene glycol monobutyl ether, and water, as summarized by Holm
oils do not have a sufficiently high acid number for this approach (1977). Slugs of these materials miscible displaced oil and with
to work. better sweep than previous solvents. However, it was not initially
Foam can be used for mobility control by alternating slugs of recognized that process success also depended on maintaining
gas with slugs of surfactant solution. Besides effective oil dis- ultralow IFT at the rear of the slug, where it was displaced by an
placement in a homogeneous sandpack, it demonstrated greatly aqueous polymer solution and became a Winsor I microemulsion
improved sweep in a layered sandpack. (Hirasaki 1981).
The other approach involved injection of a surfactant formula-
Introduction tion made of a petroleum sulfonate and alcohol in an aqueous
It is generally considered that only approximately one-third of the electrolyte solution. Key to the success of this approach were
petroleum present in known reservoirs is economically recoverable systematic studies of oil displacement leading to recognition that a
with established technology (i.e., primary-recovery methods using dimensionless capillary number N c = ␮ v /␴ controlled the amount
gas pressure and other natural forces in the reservoir, and second- of residual oil remaining after flooding an oil-containing core at
ary recovery by waterflooding). It has long been an objective of interstitial velocity v with an aqueous solution having a viscosity
the industry to develop improved processes to increase overall µ and IFT ␴ with the oil (Taber 1969; Stegemeier 1977; Melrose
recovery. However, the low oil prices that prevailed from the mid- and Brandner 1974; Foster 1973). In situations when gravity is
1980s until recently provided little incentive for research on EOR, important, the Bond number must be included (Pennell et al. 1996).
especially surfactant processes with substantial initial cost for This work revealed that at typical reservoir velocities, IFT had
chemicals. In light of the current higher prices and accompanying to be reduced from crude-oil/brine values of 20 to 30 mN/m to
values in the range of 0.001 to 0.01 mN/m to achieve low values
of residual-oil saturation (<0.05).
Several research groups found that ultralow IFTs in the required
Copyright © 2011 Society of Petroleum Engineers
range could be achieved using petroleum-sulfonate/alcohol mix-
This paper (SPE 115386) was accepted for presentation at the SPE Annual Technical tures (Hill et al. 1973; Foster 1973; Cayias et al. 1977). They
Conference and Exhibition, Denver, 21–24 September 2008, and revised for publication.
Original manuscript received for review 19 July 2010. Revised manuscript received for
also found systematic variations of IFT when changing such vari-
review 22 December 2010. Paper peer approved 18 January 2011. ables as salinity, oil composition, and temperature. An important

December 2011 SPE Journal 889


contribution was the work of Healy et al. (1976) [see also Reed Surfactant Requirements and Structures
and Healy (1977)], who demonstrated a relationship between IFT In a successful displacement process, the injected surfactant slug
and microemulsion phase behavior. Core tests using continuous must first achieve ultralow IFT to mobilize residual oil and create
surfactant injection at the optimal salinity also yielded the highest an oil bank where both oil and water flow as continuous phases
recovery of waterflood residual oil. Their studies used mixtures of (Bourrel and Schechter 1988). Second, it must maintain ultralow
an alcohol cosolvent with synthetic alkyl/aryl sulfonates, in par- IFT at the moving displacement front to prevent mobilized oil from
ticular C9, C12, and C15 orthoxylene sulfonates, which can be made being trapped by capillary forces. Because of the way they are pre-
from oligomers of propylene with more reproducible compositions pared, commercial surfactants are invariably mixtures of multiple
than those belonging to petroleum sulfonates. species, which raises questions as to whether chromatographic
separation (i.e., preferential adsorption on pore surfaces or pref-
Conventional Phase Behavior for Ultralow IFT. The under- erential partitioning into the oil phase of some species) can cause
standing of ultralow IFT in oil-recovery processes was advanced IFT variations with possible adverse effects on oil recovery. When
when Healy et al. (1976) explained how the Winsor definition of alcohol is used in the formulation, it partitions among the bulk-oil
equilibrium microemulsion phase behavior (I, II, and III, or lower- and brine phases and the surfactant films in a manner different
phase, upper-phase, and middle-phase microemulsion, respec- from the surfactant. The alcohol must then be carefully selected
tively) described the changes of phase behavior, solubilization of and tested to ensure there is no deleterious effect of chromato-
oil and water, and IFT as a function of salinity for anionic surfac- graphic separation (Dwarakanath et al. 2008; Sahni et al. 2010).
tants. The surfactant is able to solubilize an increasing amount of In the surfactant films, alcohol serves as a cosolvent, making the
oil and a decreasing amount of water as salinity is increased. The films less rigid and thereby increasing the rate of equilibration and
“optimal salinity” determined from phase behavior is the salinity preventing formation of undesirable viscous phases and emulsions
at which the microemulsion solubilizes equal amounts of oil and instead of the desired low-viscosity microemulsions. Alcohol can
water. The optimal salinity at which equilibrium IFTs between also serve as a cosurfactant, altering, for instance, the optimal salin-
the microemulsion phase and excess-oil or excess-water phase ity required to achieve ultralow IFT. Alcohols with short chains
become equal (and thus the sum becomes a minimum) is close to such as propanol increase optimal salinity for sulfonate surfactants,
the optimal salinity from phase behavior. There are correlations while longer-chain alcohols such as pentanol and hexanol decrease
between the “solubilization parameters” (ratio of oil/surfactant optimal salinity. For petroleum sulfonates and synthetic alkyl/aryl
Vo /Vs or water/surfactant Vw /Vs by volume) and the IFTs of the sulfonates with light crude oils, it has been found that 2-butanol
microemulsion with the respective excess phases (Huh 1979). acts as a cosolvent but has less effect on optimal salinity than
Thus, one can estimate the value of the equilibrium IFT at the other alcohols.
optimal salinity from the value of the solubilization parameters at A disadvantage of using alcohol is that it decreases solubiliza-
the optimal salinity (where they are equal). tion of oil and water in microemulsions, and hence increases the
Nelson and Pope (1978) recognized that the appearance of a mid- minimum value of IFT achievable with a given surfactant (Salter
dle-phase microemulsion (Winsor III) is dependent on the amounts 1977). Also, it destabilizes foam that may be desired for mobility
of water, oil, and surfactant present. Thus, they defined the Type III control with the slug and in the drive. For temperatures below
phase environment as the range of salinity at which a middle-phase approximately 60°C, the need for alcohol’s cosolvent effect can
microemulsion may exist if one were to scan the water/oil/surfactant be reduced or eliminated by some combination of the following
ternary diagram. This distinction is important at very high or very low strategies: (1) using surfactants with branched hydrocarbon chains,
surfactant concentrations because the volume of the middle-phase (2) adding ethylene oxide (EO) and/or less-hydrophilic propylene
microemulsion is proportional to the surfactant concentration. At high oxide (PO) groups to the surfactant, and (3) using mixtures of
surfactant concentrations, more of the excess phases are solubilized, surfactants with different hydrocarbon-chain lengths or structures.
and thus the excess phases have smaller volume or are not present. Such measures counter the tendency of long, straight hydrocarbon
If the surfactant concentration is high enough, the “middle-phase” chains of nearly equal length to form condensed surfactant films
microemulsion phase may appear as a single phase at or near optimal and the lamellar liquid-crystalline phase. At high temperatures,
conditions. On the other hand, at low surfactant concentrations but increased thermal motion promotes more-flexible surfactant films
above the critical micelle concentration, the volume of the middle- and disruption of ordered structures, but it does not always elimi-
phase microemulsion is minute and its presence may not be visually nate viscous phases and emulsions. Further studies are needed
detected or sampled for IFT measurements. to investigate usefulness of these or other strategies for high
The nanostructure of the microemulsion should be recognized temperatures.
to distinguish it from macroemulsions or liquid-crystal dispersions The use of branched hydrocarbon chains to minimize or elimi-
or phases. Macroemulsions are nonequilibrium dispersions that nate alcohol requirements was discussed by Wade, Schechter, and
change with time or may be in a metastable condition. Liquid- coworkers (Abe et al. 1986). An isotridecyl hydrophobe was used
crystal phases are condensed, ordered phases that usually are in Exxon’s pilot test in the Loudon field. The hydrophilic part of
birefringent (rotate polarized light), viscous, and tend to inhibit the surfactant was a chain consisting of short PO and EO segments
emulsion coalescence (Healey and Reed 1974). Microemulsions and a sulfate group (Maerker and Gale 1992). With this combina-
are equilibrium isotropic phases that may have a bicontinuous tion of the first two strategies, no alcohol was required. The Neodol
structure with near-zero mean curvature at or near optimal condi- 67 hydrophobe developed and manufactured by Shell Chemical has
tions (Scriven 1976). (Microemulsions are oil-swollen micelles in an average of 1.5 methyl groups added randomly along a straight
water at underoptimum conditions and reversed micelles in oil at C15–C16 chain that provides another type of branching. A propoxyl-
overoptimum conditions.) It was once thought that it is necessary ated sulfate with this hydrophobe has been blended with an internal
to have a cosolvent (alcohol) to have a microemulsion with an olefin sulfonate, which is also branched, and used to displace west
anionic surfactant. However, it is now recognized that it is possible Texas crude oils in low-salinity, ambient-temperature laboratory
to have microemulsions without alcohol at room temperature by tests, both with (Levitt et al. 2009) and without (Liu et al. 2008)
using branched surfactants (Abe et al. 1986). alcohol. In this case, all three strategies were combined.
Salinity-scan tests are used routinely to screen phase behavior Long-term surfactant stability at reservoir conditions is another
of surfactant formulations before conducting more time-consuming surfactant requirement. Provided that pH is maintained at slightly
coreflood tests (Levitt et al. 2009; Flaaten et al. 2009; Mohammadi alkaline levels and calcium concentration is not too high, hydro-
et al. 2009). The minimum IFT is correlated with the solubilization lysis of sulfate surfactants is limited for temperatures up to
parameters at the optimal salinity. The presence of viscous, struc- 50–60°C (Talley 1988). Surfactants with other head groups, most
tured, or birefringent phases and/or stable macroemulsions is easily likely sulfonates or carboxylates, will be needed for reservoirs
observed. Apparent viscosities of phases present in 5-mL samples at higher temperatures. Because the sulfonate group is added at
in sealed glass pipettes can be measured by the falling-sphere different points along the hydrocarbon chain during synthesis,
method, even for opaque phases (Lopez-Salinas et al. 2009). internal olefin sulfonate (IOS) surfactants consist of species with

890 December 2011 SPE Journal


a twin-tailed structure, a different type of branching. Results of substantial volumes of crude oil in the reservoir (Maerker and
laboratory studies of IOS phase behavior at high temperatures Gale 1992).
have recently been presented (Barnes et al. 2008, 2010; Zhao As temperature increases, the lamellar liquid-crystal phase
et al. 2008, Puerto et al. 2010). Hydrocarbon-chain lengths ranged may melt. However, the surfactant/brine mixture is still unsuitable
from C15–C18 to C24–C28. However, the effect of dissolved calcium for injection if separation into two or more liquid phases occurs
and magnesium ions, which most likely cause surfactant precipita- near optimal salinity (Benton and Miller 1983). Even if bulk
tion, was not investigated in these studies (see Fig. 3 of Liu et al. phase separation does not occur, turbid solutions are sometimes
2008). Moreover, the current availability of internal olefins of long observed. These solutions usually have large, anisotropic micelles
chain length, as would be required for low or moderate reservoir and separate into surfactant-rich and polymer-rich phases with the
salinities, is limited. Alpha olefin sulfonate (AOS) surfactants have addition of polymer. These phases can separate and/or plug the
the same potential disadvantages and are not as highly branched porous media into which they are injected. Adding alcohol can
as IOS surfactants. reduce micelle size and prevent phase separation in some cases.
Sulfonate groups cannot be added directly to alcohols, includ- As indicated previously, addition of a paraffinic-oil which yields an
ing those with EO and/or PO chains. One approach is to prepare oil-in-water microemulsion with nearly spherical drops is another
sulfonates with glycidyl chloride or epichlorohydrin, where a approach for preventing phase separation when polymer is present.
three-carbon chain is added between the EO or PO chain and Within limits, the higher the molecular weight of the oil added to
the sulfonate group. Wellington and Richardson (1997) described produce an oil-in-water microemulsion, the less oil is needed to
some results with such surfactants. Further information on their formulate single phases with polymer for mobility control.
synthesis and initial phase-behavior results for propoxy glycidyl Another approach to formulate single-phase injection composi-
sulfonates with the Neodol 67 hydrophobe is given by Barnes tions would be to find surfactants or surfactant blends that neither
et al. (2008). Puerto et al. (2010) presented phase behavior showing exhibit phase separation nor form turbid solutions or liquid crystal-
that suitable ethoxy or propoxy glycidyl sulfonates could produce line dispersions at conditions of interest (Flaaten et al. 2009; Sahni
microemulsions with high solubilization over a wide range of opti- et al. 2010). Blends of the branched surfactant Neodol 67 propoxyl-
mal salinities with n-octane as the oil at temperatures up to 120°C. ated sulfate (N67-7POS), having an average of seven PO groups
They also noted that because these surfactants typically exhibit with the twin-tailed surfactant IOS 15-18, an IOS (Barnes et al.
phase separation (cloud point) at high temperatures and salinities, 2010) made from a feedstock containing mainly C15–C18 chains, are
it may be necessary to blend them with other surfactants such as interesting in this respect. Fig. 1 shows phase behavior at ambient
IOS, which are also stable at high temperatures. temperature of 3 wt% aqueous solutions of such surfactant blends
Until recent years, nearly all work was directed toward EOR in containing 1 wt% Na2CO3 and varying NaCl concentration but no
sandstone reservoirs owing to concerns that in the high-divalent- alcohol or polymer. IOS 15-18 alone precipitates above 4 wt%
ion environment of carbonate reservoirs, petroleum or synthetic NaCl in such solutions. In contrast, solutions of the propoxylated
sulfonates would adsorb excessively and/or form calcium and sulfate alone do not precipitate but instead become cloudy above
magnesium salts that would either precipitate or partition into the the same salinity, as droplets of a second liquid phase form and
oil phase. The exception was the work of Adams and Schievelbein scatter light. Addition of IOS 15-18 to the propoxylated sulfate
(1987), who conducted laboratory experiments and two field tests makes the mixture more hydrophilic, thereby raising the salinity at
showing that oil could be displaced in a carbonate reservoir using which phase separation occurs to a value higher than for either sur-
a mixture of petroleum sulfonates and ethoxylated sulfate surfac- factant alone. For instance, the 4:1 blend (hereafter NI blend) (i.e.,
tants. The ethoxy groups add tolerance to divalent ions. Recent 80% N67-7POS and 20% IOS 15-18) exhibits phase separation
work with carbonate reservoirs used ethoxylated or propoxylated at approximately 6 wt% NaCl (plus 1% Na2CO3), although slight
sulfates, as discussed in the later section on wettability. cloudiness occurs above approximately 3.5 wt% NaCl. Addition
of 0.5 wt% of partially hydrolyzed polyacrylamide to a 0.5 wt%
solution of this blend in a solution containing 4 wt% NaCl and 1
Alcohol-Free Surfactant Slugs for Injection wt% Na2CO3 produces phase separation, although similar addition
The surfactant slug to be injected should be a single-phase micel- of polymer to a solution containing only 2 wt% NaCl does not
lar solution. Especially when polymer is added to increase slug (Liu et al. 2008). Phase-behavior studies show that the optimal
viscosity, it is essential to prevent separation into polymer-rich salinity of this blend with a west Texas crude oil is approximately
and surfactant-rich phases, which yields highly viscous phases 5 wt% NaCl (with 1 wt% Na2CO3) when the amount of surfactant
unsuitable for either injection or propagation through the forma- present is much greater than soap formed from the naphthenic
tion (Trushenski 1977). At low temperatures, oil-free mixtures of acids in the crude oil. However, in alkaline/surfactant processes, it
petroleum sulfonate/alcohol or synthetic sulfonate/alcohol mix- is best to inject at lower salinities, as discussed later, because the
tures with brine are often translucent micellar solutions at salinities surfactant encounters conditions during the process with greater
well below optimal, but contain lamellar liquid crystal and exhibit ratios of soap-to-surfactant and correspondingly lower optimal
birefringence near optimal salinity where ultralow IFT is found salinities. Indeed, excellent recovery of the west Texas crude oil
upon mixing with crude oil (Miller et al. 1986). In the absence was observed in sandpack experiments when a single-phase mix-
of polymer, the lamellar phase is often dispersed in brine as ture of the 4:1 blend and polymer was injected at 2 wt% NaCl
particles having maximum dimensions of at least several microm- with 1 wt% Na2CO3 (Liu et al. 2008). At 2% NaCl, the surfactant
eters. When polymer is added to such a turbid dispersion of the micelles are not highly anisotropic, and polymer and surfactant can
lamellar phase, it produces a polymer-rich aqueous solution and a coexist in the same phase.
more concentrated surfactant dispersion (Qutubuddin et al. 1985). A similar approach was used by Falls et al. (1994) in an
This undesirable behavior can sometimes be avoided by adding alkaline/surfactant field test. They added a small amount of the
sufficient alcohol. However, use of alcohol has disadvantages, as nonionic surfactant Neodol 25-12 to the main injected surfactant, a
indicated previously. blend of IOSs, to make the formulation sufficiently hydrophilic to
The lamellar phase was observed in surfactant/brine mixtures in form a single micellar solution during storage at ambient tempera-
the absence of oil even for Exxon’s Loudon formulation mentioned ture. Because this surfactant becomes less hydrophilic at higher
previously (Ghosh 1985), where, as indicated, branching and temperatures, it did not adversely affect process performance at
addition of EO/PO groups allowed low-viscosity microemulsions the reservoir temperature of approximately 57°C.
to be formed and ultralow IFT to be achieved with the crude oil
without the need to add alcohol. Exxon avoided phase separation Alkaline/Surfactant Processes: Role of Alkali
when polymer was added to the injected slug by including a paraf- Nelson et al. (1984) proposed injection of a solution containing
finic white oil of high molecular weight in the formulation. That both surfactant and alkali for EOR. Such processes have attracted
is, they injected a white oil-in-water microemulsion (Winsor I), and continue to attract considerable interest. They have been
which became a bicontinuous microemulsion when mixed with labeled by different names, but will be collectively described here

December 2011 SPE Journal 891


10

9 Multiphase
Region
8 Phase boundary
7 Clear solution
6 Two clear phases
% NaCl
5 Precipitation
4 One-Phase
Region Cloudy solution
3
Cloudy after nine
2 months

0
1:1 4:1 9:1
IOS N67
N67:IOS (w/w)

Fig. 1—Effect of added NaCl on phase behavior of 3 wt% solutions of N67/IOS mixtures containing 1 wt% Na2CO3 (Liu et al.
2008).

as alkaline/surfactant processes (Nelson et al. 1984; Peru and ponent of surfactant retention. Phase trapping of surfactant can be
Lorenz 1990; Surkalo 1990; Baviere et al. 1995). more significant and will be discussed later.
The primary role of the alkali in an alkaline/surfactant process Alkaline preflush had been advocated for both sequestering
is to reduce adsorption of the surfactant during displacement divalent ions and reducing sulfonate adsorption (Holm and Rob-
through the formation and sequestering divalent ions. An addi- ertson 1981). In subsequent work, alkali has been injected with the
tional benefit of alkali is that the soap is formed in situ from the surfactant. Adsorption of anionic surfactants on Berea sandstone
naphthenic acid in the crude oil (Johnson 1976). As indicated pre- was reduced several-fold with addition of sodium carbonate for
viously, the generation of soap allows the surfactant to be injected petroleum sulfonate (Bae and Petrick 1977) or with addition of
at lower salinities than if used alone, which further reduces adsorp- sodium silicate or hydroxide for alcohol ethoxysulfate (Nelson
tion and facilitates incorporation of polymer in the surfactant slug. et al. 1984). The reduction of adsorption on Berea sandstone with
Also, alkali can alter formation wettability to reach either more sodium bicarbonate was 68% in a dynamic experiment (Peru and
water-wet or more oil-wet states. In fractured oil-wet reservoirs, Lorenz 1990). Static and dynamic adsorption of anionic surfactants
the combined effect of alkali and surfactant in making the matrix on calcite and dolomite was decreased by an order of magnitude
preferentially water-wet is essential for an effective process. These with addition of sodium carbonate, but insignificantly with sodium
benefits of alkali will occur only where alkali is present. Thus, it hydroxide (see Figs. 2 through 4) (Hirasaki and Zhang 2004;
is important to determine “alkali consumption,” which controls the Seethepalli et al. 2004; Zhang et al. 2006; Liu et al. 2008; Tabata-
rate of propagation of alkali through the formation. bal et al. 1993). (The TC Blend of Figs. 2 through 4 is an earlier
blend of isotridecyl 4PO sulfate and C12 3EO sulfate. Research
Reduced Surfactant Adsorption. The discussion here will be on it was discontinued because its optimal salinity was too high
limited to anionic surfactants (Wessen and Harwell 2000). The for the application of interest.)
primary mechanism for the adsorption of anionic surfactants on
sandstone- and carbonate-formation material is the ionic attrac- Divalent-Ion Sequestration. The phase behavior of anionic sur-
tion between positively charged mineral sites and the negative factant systems is much more sensitive to a change in divalent ions
surfactant anion (Tabatabal et al. 1993; Zhang and Somasundaran (e.g., Ca2+ and Mg2+) compared to monovalent ions (e.g., Na+),
2006). Thus, the role of the alkali is to be a “potential-determining especially at low surfactant concentrations (Nelson 1981). This is
ion” to reverse the charge on positively charged mineral sites. The problematic in sandstones because of ion exchange between the
potential-determining ions for oxide minerals are the hydronium clay, brine, and surfactant micelles (Hill et al. 1977; Pope et al.
and hydroxide ions. The pH at which the charge reverses is the 1978; Hirasaki 1982). This exchange can result in the phase behav-
“isoelectric point” if measured by electrophoresis (zeta potential) ior becoming overoptimum, with resulting large surfactant reten-
and is the “point-of-zero-charge” if determined by titration. The tion (Glover et al. 1979, Gupta 1982). Alkali anions (e.g., carbon-
values are tabulated for most common minerals (Lyklema 1995). ate, silicate, and phosphate) that have low solubility product with
Silica is negatively charged at reservoir conditions and exhibits divalent cations will sequester divalent cations to low concentra-
negligible adsorption of anionic surfactants. Clays (at neutral pH) tions (Holm and Robertson 1981). Hydroxide is not as effective for
have negative charge at the faces and positive charge at the edges. sequestration of calcium because the solubility product of calcium
The clay edges are alumina-like and thus are expected to reverse hydroxide is not very low. Sodium metaborate has recently been
their charge at a pH of approximately 9. Carbonate formations and introduced as an alkali that may sequester divalent ions (Flaaten
sandstone-cementing material can be calcite or dolomite. These et al. 2009; Zhang et al. 2008). A common problem with alkali
latter minerals also have an isoelectric point of approximately pH injection is that softened water is needed to avoid scaling.
9, but carbonate ions, as well as the calcium and magnesium ions,
are more significant potential-determining ions. The zeta potential Generation of Soap. The original concept of alkali flooding was
of calcite is negative even at neutral pH in the presence of 0.1 N the reduction of oil/water IFT by in-situ generation of soap, which
carbonate/bicarbonate ions (Hirasaki and Zhang 2004). If a for- is an anionic surfactant, sodium naphthenate (Jennings 1975).
mation contains iron minerals, the oxidation/reduction conditions Ultralow IFT usually required injection of relatively fresh water
influence whether the surface iron sites are Fe3+ or Fe2+. Adsorption with a low concentration of alkali because optimal salinity (total
of anionic surfactant for one sandstone was found to be lower by electrolyte concentration) of the in-situ-generated soap is usually
more than a factor of two for reducing rather than for oxidizing low (e.g., <1% electrolyte). If the alkali concentration is too low,
conditions (Wang 1993). Surfactant adsorption is only one com- alkali consumption reactions may result in a large retardation of

892 December 2011 SPE Journal


2.5

2.0

Adsorption Density,
10–3 mmol/m2
1.5

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5

Residual Surf. Conc., mmol/l

Original 0.05%TC blend 0.05%TC blend 0.05%TC blend/3% Na2CO3


0.05%TC blend/3.5% Na2CO3 0.05%TC blend/4% Na2CO3 0.05%TC blend/4.5% Na2CO3
Original 0.1%TC blend 0.1%TC blend 0.1%TC blend//3% Na2CO3

Fig. 2—Static adsorption of TC blend surfactant on dolomite sand. BET surface area of the calcite: 17.8 m2/g (Zhang et al.
2006).

the alkali displacement front. The concept of alkaline/surfactant Alkali Consumption. The ASP process should be designed such
flooding is to inject a surfactant with the alkaline solution such that that displacement fronts of the alkali, surfactant, and polymer
mixture of the in-situ-generated soap and injected surfactant has travel together. The mechanisms responsible for the retardation of
an optimal salinity that is tailored to the reservoir fluids (Nelson the alkali front include silica dissolution, clay dissolution with zeo-
et al. 1984; Surkalo 1990). lite precipitation, anhydrite or gypsum dissolution with calcite (or
The common method used to determine the amount of naph- calcium hydroxide or silicate) precipitation, dolomite dissolution
thenic acid in crude oil is the total acid number (TAN), determined with calcium and magnesium silicate precipitation, hydrogen-ion
by nonaqueous titration with a base (Fan and Buckley 2007). If exchange, divalent-ion exchange with precipitation, and mixing
sodium naphthenate is to act as a surfactant, it should partition with divalent ions in formation water with precipitation (Ehrlich
into the aqueous phase at low electrolyte concentrations and be and Wygal 1977; Holm and Robertson 1981; Southwick 1985;
measurable by hyamine titration for anionic surfactants. It was Cheng 1986; Novosad and Novosad 1984; Jensen and Radke 1988;
found that the sodium naphthenate determined by extraction into Mohammadi et al. 2009). Naphthenic acids in crude oil also react
the aqueous phase and measured by hyamine titration is less than with alkali and thus contribute to consumption, but the amount
the TAN value (Liu et al. 2010). It is hypothesized that the TAN is usually small compared to the mentioned inorganic mineral
includes components that are too lipophilic to be extracted to the reactions. Silica dissolution can be controlled by using a buffered
aqueous phase and/or too hydrophilic to be detected by hyamine system such as sodium carbonate or silicate rather than hydroxide
titration. (Southwick 1985). Clay dissolution is strongly dependent on the

1.0
Expr. 1
0.9 v=1.2 feet/day
beta=0.34±0.03 Expr. 2
Cl−
0.8 Φ=0.335
Dimensionless Concentration

0.7 Expr. 2
Expr. 1
v=12 feet/day
0.6 beta=0.22±0.03
Φ=0.338
0.5
Calculated
0.4 from dolomite
isotherm
0.3
beta=0.04
0.2

0.1

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Injected Volume, PV

Fig. 3—Dynamic adsorption of 0.2% TC blend surfactant without Na2CO3 on dolomite sand (Zhang et al. 2006).

December 2011 SPE Journal 893


1.0
0.1%CS330+0.1%TDA-4PO
0.9 +0.3MNa 2 CO 3
0.8 beta=0.07±0.04

Dimensionless Concentration
v=1.2 feet/day
0.7

0.6

0.5
Experimental Data for NaCl
0.4

0.3 Experimental Data for Surfactant

0.2 Simulation Curve for Surfactant


0.1
Simulation Curve for NaCl
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Injected Volume, PV

Fig. 4—Dynamic adsorption of 0.2% TC blend/0.3-M Na2CO3 on dolomite sand (Zhang et al. 2006).

pH and type of clay, and is kinetically limited (Sydansk 1982). be preferentially water-wet when the salinity is below the optimal
Thus, acidic clay such as kaolinite, as well as high temperature, salinity (Winsor I). When the system is overoptimum (Winsor
will increase the importance of this mechanism. II), macroemulsions tend to be oil-external. An oil-external mac-
A limitation of the application of sodium carbonate in carbonate roemulsion will trap water and have a low oil and water relative
formations is that if anhydrite or gypsum is present, it will dissolve permeability, similar to what one expects with oil-wet porous
and precipitate as calcite (Hirasaki et al. 2005; Liu 2007). This is media. The optimal salinity for a conventional alkali flooding
detrimental for dolomite formations because they may have origi- system is dependent on the in-situ-generated sodium naphthenate
nated from evaporite deposits where gypsum is usually present. An soap, and is usually below approximately 1% electrolyte strength.
alternative alkali is sodium metaborate (Zhang et al. 2008; Flaaten Because salinity of reservoir brine typically exceeds this value, a
et al. 2009). However, longer-term experiments and equilibrium conventional alkali flood often generates overoptimum and oil-wet
calculations indicate that this metaborate will also precipitate. conditions. We show later in this review that this behavior can
be avoided by injecting the alkali and surfactant in the Winsor I
Alkaline/Surfactant Processes: Wettability region. After mixing with the fluids in the reservoir of interest, it
Alteration will pass through the Winsor III, low-IFT region. Even a high-
Wettability is the next most important factor in waterflood recov- salinity sandstone formation that is initially oil-wet may be altered
ery after geology (Morrow 1990). The recovery efficiency of a to preferentially water-wet by injecting alkali with a hydrophilic
flooding process is a function of the displacement efficiency and surfactant in the Winsor I region.
sweep efficiency. These efficiencies are a function of the residual-
oil saturation (waterflood and chemical flood) and mobility ratio, Carbonate Formations. Wettability alteration has received more
respectively. The residual-oil saturation to waterflooding is a attention recently for carbonate formations compared to sand-
function of wettability, with the lowest value at intermediate wet- stones because carbonate formations are much more likely to be
tability (Jadhunandan and Morrow 1995). The mobility ratio is a preferentially oil-wet (Treiber and Owens 1972). Also, carbonate
function of the ratio of water relative permeability to oil relative formations are more likely to be fractured and will depend on
permeability at their respective endpoints or at a specific satura- spontaneous imbibition or buoyancy for displacement of oil from
tion. The mobility ratio or relative permeability ratio becomes the matrix to the fracture.
progressively larger as the wettability changes from water-wet to Wettability-alteration tests on plates of calcite, marble, lime-
oil-wet (Anderson 1987b). When a formation is strongly oil-wet, stone, and dolomite with different surfactants and sodium carbon-
it can have both a high waterflood residual-oil saturation and ate have been used to identify many systems that are altered to
unfavorable mobility ratio. In addition, an oil-wet formation will preferentially water-wet with low anionic-surfactant concentra-
have capillary resistance to imbibition of water (Anderson 1987a). tions (Hirasaki and Zhang 2004; Seethepalli et al. 2004; Zhang
Formation wettability can be altered by pH (Wagner and Leach et al. 2006; Adibhatla and Mohanty 2008; Gupta et al. 2009).
1959; Ehrlich et al. 1974; Takamura and Chow 1985; Buckley Sodium carbonate has an important role because the carbonate ion
et al. 1989, Dubey and Doe 1993), surfactants that adsorb on the is a potential-determining ion for calcite and dolomite (Hirasaki
minerals (Somasundaran and Zhang 2006) or remove adsorbed and Zhang 2004).
naphthenic acids (Standnes and Austad 2000), and acids or bases
(Cuiec 1977). These processes are now incorporated into chemi- Spontaneous Imbibition. Spontaneous imbibition is the process
cal-flood simulators (Anderson et al. 2006; Delshad et al. 2006; by which a wetting fluid is drawn into a porous medium by capil-
Adibhatla and Mohanty 2007, 2008). lary action (Morrow and Mason 2001). The presence of surfactant
in some cases lowers the IFT, and thus the capillary pressure, to
Sandstone Formations. Wettability alteration to more water-wet negligible values. Spontaneous displacement can still occur in this
or more oil-wet conditions was proposed as one of the mechanisms case by buoyancy or gravity drainage (Schechter et al. 1994).
of caustic flooding (Wagner and Leach 1959; Ehrlich et al. 1974; The research group of Austad has investigated spontane-
Johnson 1976). Our current understanding of microemulsion phase ous imbibition into chalk-formation material with enhancement
behavior and wettability is that the system wettability is likely to by cationic and nonionic surfactants and/or sulfate ions present in

894 December 2011 SPE Journal


seawater (Austad et al. 1998; Standnes and Austad 2000; Milter III or middle-phase microemulsion to have ultralow IFT during
and Austad 1996a, 1996b; Høgnesen et al. 2004, 2006). Sponta- the oil-recovery displacement process. Recently, it was found that
neous imbibition was most effective with dodecyl trimethylam- blending alkyloxylated sulfate and IOS surfactants in an alkaline
monium or amine surfactants. The mechanism is thought to be environment produced ultralow IFT for a wide range of conditions
removal of adsorbed naphthenic acids through ion pairing with the even when the system has Winsor I phase behavior. This is an
cationic surfactant. Capillarity may dominate during earlier time, important discovery because surfactant retention is much less for
with gravity dominating later. Winsor I compared with Winsor III phase behavior. These results
Laboratory and field testing of surfactant-enhanced imbibition are reviewed here.
was investigated for the dolomite formation of the Yates field
(Chen et al. 2000; Yang and Wadleigh 2000). Both mass-balance Middle Layer of Lower-Phase Microemulsion. When a salin-
and CT scans showed increased oil recovery with 0.35% nonionic ity scan test is conducted at low surfactant concentrations (e.g.,
and ethoxylated sulfate surfactants compared to reservoir brine. 0.05%), the equilibrium phase behavior appears to go from a
A single-well injection, soak, and production test showed increased lower-phase microemulsion to an upper-phase microemulsion over
oil recovery with decrease in water/oil ratio (WOR). The CT scans a narrow salinity range (Fig. 6) (Zhang et al. 2006: Liu et al.
showed that with formation brine, the oil recovery involved coun- 2008). Middle-phase microemulsions are rarely seen at low sur-
tercurrent imbibition, but with surfactant, the displacement was factant concentrations; thus, IFT measurements for salinity scans
dominated by gravity after early time. for low surfactant concentration are usually between the upper
An investigation compared the spontaneous recovery of the and lower phases observed in the sample tubes. The phases may
Yates system with 0.35% ethoxylated alcohol and dodecyl tri- be (1) lower-phase microemulsion and excess oil (Winsor I), (2)
methylammonium bromide [C12TAB, Standnes et al. (2002)]. excess brine and upper-phase microemulsion (Winsor II), or (3)
The superior recovery of C12TAB was interpreted to be caused by excess brine and excess oil (Winsor III). The value of the IFTs
advancing contact angle of 32° for C12TAB, compared with 107° for these systems, measured between the upper and lower phases,
for the ethoxylated alcohol and 133° for brine. was not reproducible until a protocol was developed to include a
Laboratory measurements of spontaneous oil recovery were small volume of intermediate-density material in conducting spin-
made on the Yates system or outcrop limestone with alkaline/ ning-drop measurement. In Winsor III, this material would include
surfactant solutions with 0.05% anionic surfactant and sodium the middle-phase microemulsion. In Winsor I (lower-phase micro-
carbonate (Hirasaki and Zhang 2004; Seethepalli 2004; Zhang emulsion or underoptimum salinity), this intermediate-density
et al. 2006; Adibhatla and Mohanty 2008). There was no recovery material appeared to be a colloidal dispersion in the lower-phase
(Yates cores) or only little recovery (outcrop cores) with brine microemulsion (Fig. 7). This colloidal dispersion formed a middle
as the imbibing fluid. However, the alkaline/surfactant solutions layer between the excess oil and lower-phase microemulsion of
recovered as much as 60% of the oil. For a given system, tem- alkaline/surfactant systems with crude oil with a TAN of 0.3 mg
perature is an important factor for wettability alteration and rate KOH/g or greater. This middle layer was more opaque than the
of imbibition oil recovery (Gupta and Mohanty 2010). rest of the lower-phase microemulsion, and can be interpreted to
Surfactant-aided wettability alteration and spontaneous oil be more oil-rich from its intermediate density. When the oil/water
recovery may not have a significant contribution from capillary ratio is increased, the volume of this layer also increased. Because
pressure if the IFT is reduced to ultralow values. However, in these this layer is not observed in the absence of alkali, it is hypothesized
conditions, gravity becomes an important contribution (Babadagli that the dispersed colloidal material is an oil-and-sodium naphthe-
2001; Hirasaki and Zhang 2004, Zhang et al. 2006; Adibhatla and nate-rich microemulsion that is in equilibrium with the remainder
Mohanty 2007; Gupta and Mohanty 2010, 2008). Fig. 5 illustrates of the lower-phase microemulsion. As shown later, the soap is
the simulated velocity field during spontaneous displacement for more lipophilic than the added surfactant. This layer is not just a
a cylindrical core immersed in a 0.05% alkaline/surfactant solu- macroemulsion dispersion that has not yet coalesced because the
tion. The aqueous phase enters from the sides, and oil flows out presence of this material affects the value of the IFT.
from the top of the core. The significance of gravity-dominated
displacement is that time scales proportionally with the character- IFT With and Without Alkali. The IFT of the system with and
istic length of the matrix, whereas it would scale with the square without alkali was measured to test the hypothesis that the soap
of the length if the process were capillary-dominated (Hirasaki generated by the alkali is responsible for the ultralow IFT in the
and Zhang 2004). presence of the colloidal-dispersion material. Fig. 8 compares
Najafabadi et al. (2008) and Delshad et al. (2009) simulated the measured IFT with and without alkali. In the absence of
cases of fractured oil-wet systems where negative capillary pres- alkali, the lower-phase microemulsion was homogeneous, and the
sure inhibited oil displacement from the matrix. Wettability altera- ultralow IFT occurred only near the optimal salinity, as expected
tion and low IFT reduced the unfavorable capillary pressure so that for conventional surfactant EOR systems. In the presence of
the pressure gradient from flow in the fractures could displace oil alkali and using the protocol to ensure that a small volume of
from the matrix by viscous forces. colloidal dispersion was present, a wider, ultralow IFT region was
observed, especially for underoptimum conditions. If the colloidal
Emulsification of Heavy Oil. Alkali and surfactant can be used to dispersion is not present as a result of creaming or centrifugation,
recover viscous, heavy oil by emulsification and wettability altera- IFT behavior is similar to that in the absence of alkali [Liu et al.
tion along with displacing the oil as a lower-viscosity, oil-in-water (2008), Fig. 14]. Apparently, this colloidal dispersion contains
emulsion (Liu et al. 2006; Bryan et al. 2008; Li et al. 2010; Kumar surface-active species responsible for lowering IFT between the
and Mohanty 2010; Kumar et al. 2010). lower-phase microemulsion and the excess-oil phase in a manner
similar to behavior of the middle-phase microemulsion between
Alkaline/Surfactant Processes: Wide Region the excess-brine and excess-oil phases.
of Ultralow IFT
Surfactant processes for enhancing oil recovery are based on Consistent With Solubilization Parameter. The wide range of
achieving ultralow IFT (e.g., <10–2 mN/m) to either raise the capil- ultralow IFT was verified by comparison with the Huh correlation
lary number in forced displacement or to raise the Bond number (Huh 1979). The solubilization ratios of the alkaline NI blend with
for gravity-driven displacement (Pennell et al. 1996) from a matrix Yates crude oil are shown in Fig. 9. The IFTs calculated from
that is surrounded by fractures or vugs bearing surfactant solution. the Huh correlation and measured by spinning drop are shown
For most anionic surfactants, especially those sensitive to divalent in Fig. 10. Thus, the measured IFT is consistent with the volume
ions, ultralow IFT occurs only over a limited range of conditions of oil that is solubilized into the microemulsion phase(s)/volume
[Healy et al. (1976); IFT vs. salinity plot shows a narrow region of surfactant. The generality of the wide ultralow IFT was tested
of ultralow IFT]. It has always been a process-design challenge with another crude oil with a TAN of 4.79 mg KOH/g. The lower-
to maintain or pass through these narrow conditions of Winsor phase microemulsion of this crude oil was too dark to observe the

December 2011 SPE Journal 895


Aqueous Phase Velocity 10.8852

0.09

0.08

Distance From Bottom, m ---> 0.07

0.06

0.05

0.04

0.03

0.02

0.01

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018
(a) Distance From Center, m --->

Oil Phase Velocity 10.8852

0.09

0.08

0.07
Distance From Bottom, m --->

0.06

0.05

0.04

0.03

0.02

0.01

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018
(b) Distance From Center, m --->

Fig. 5—Velocity field in a cylindrical core immersed in surfactant solution at 10.89 days: (a) aqueous phase, (b) oil phase (Gupta
and Mohanty 2010).

spinning drop for IFT measurements. The IFT estimated from the the displacement process. Thus, an alkaline/surfactant system
Huh correlation shows this crude oil to also have a wider, ultralow should be considered as a pseudo-two-surfactant system featuring
IFT region of salinities, especially in the underoptimum region the injected surfactant and the soap. The two surfactants will likely
(Liu et al. 2010). have different optimal salinities. Thus, a mixing rule is needed to
model how the optimal salinity changes with surfactant and soap
Alkaline/Surfactant Processes: Phase concentrations.
Behavior of Soap/Surfactant
Alkali saponifies the naphthenic acid in crude oil in situ to generate WOR and Surfactant Concentration. Optimal salinity was
sodium naphthenate, a soap that helps to generate low IFT during observed to be a function of surfactant concentration and WOR

896 December 2011 SPE Journal


x= 0.2 0.8 1.4 2.0 2.6 3.2 3.6 4.0 4.5 5.0

Excess
Oil

Colloidal
Dispersion

Lower-Phase
Microemulsion

Fig. 6—Salinity scan for 0.2% NI blend, 1% Na2CO3 with MY4


crude oil for WOR = 3 after settling time of 28 days at 25°C.
x = wt% NaCl (from Liu et al. 2008).

Fig. 7—View of colloidal dispersion region near interface for


for an alkaline/surfactant system (Fig. 11). However, all of these 2% NaCl sample from salinity scan, after 23 days settling (from
curves can be reduced to a single curve if plotted as a function of Liu et al. 2008).
the soap/surfactant ratio (Fig. 12). The latter figure compares the
curves of optimal salinity for the TC blend and NI blend surfactant (An alternative to analytical determination of the extractable
formulations and the same crude oil. soap content is to estimate a value that will result in the best fit
to Eq. 1.)
Mixing Rule. The modeling of alkaline/surfactant flooding will The Salager et al. (1979) mixing rule was found to be followed
benefit from a mixing rule for the optimal salinity. When the TAN reasonably well when the aqueous-titration method was used to
was used for the soap content, the experimental data deviated quantify the soap content of the crude oil (Fig. 13, right panel).
significantly from the mixing rule of Salager et al. (1979). An The expression for the mixing rule is
alternative approach to determine the soap content of crude oil is
to extract soap from the crude oil into alkaline, alcoholic water
and titrate for anionic-surfactant content by hyamine titration. ( ) ( )
log (Optmix ) = X soap log Optsoap + 1 − X soap log (Optsurfactant )

1.E+01
Without Na2CO3

With 1% Na2CO3
1.E+00

1.E-01
IFT, mN/m

1.E-02

1.E-03

1.E-04
0 1 2 3 4 5 6
Salinity, % NaCl

Fig. 8—Measured IFT of system with and without Na2CO3 (from Liu et al. 2008).

December 2011 SPE Journal 897


1000 1.E-01
Chun-Huh Correlation

Vw /Vs Spinning Drop Measurement


Solubilization Ratio

100 1.E-02

IFT, mN/m
10 1.E-03
Vo/Vs

1 1.E-04
2 2.5 3 3.5 4 2.0 2.5 3.0 3.5 4.0
NaCl, % NaCl, %

Fig. 9—Measured solubilization ratios of salinity scan (from Fig. 10—Comparison of the IFT from the solubilization param-
Liu et al. 2008). eter and spinning-drop measurements (from Liu et al. 2008).

14 14
WOR=1 (TC blend)

Optimal NaCl Conc., %


12 12
Optimal NaCl Conc., %

WOR=3 (TC blend)


10 10 WOR=10 (TC blend)
8 8 NI blend
NI blend
6
6 TC blend
4
4 WOR=10
2
2 WOR=3
WOR=1 0
0 1.E-02 1.E-01 1.E+00 1.E+01
0.01 0.1 1 10
Soap/Synthetic Surfactant Mole Ratio
Surfactant Concentration, %
Fig. 12—Optimal salinity as a function of soap/surfactant ratio
Fig. 11—Optimal sodium chloride concentration of TC blend for NI and TC surfactant blends with MY4 crude oil (Liu et al.
as a function of WOR and surfactant concentration (settled for 2008).
more than 6 months) (from Zhang et al. 2006).

where ␴ *M = X soap ␴ soap


*
+ X surfactant ␴ surfactant
*
. . . . . . . . . . . . . . . . . . . . . (2)

Soap where ␴ *M , ␴ soap


*
, and ␴ surfactant
*
X soap = , mole fraction . . . . . . . . . . . . . . (1) are the optimal solubilization ratio of
Soap + Surfactant the mixture, soap, and surfactant, respectively.

The optimal-salinity mixing rule is used in UTCHEM for simu- IFT Measurements. The dependence of the optimal salinity on
lation of alkaline/surfactant processes (Mohammadi et al. 2009). the soap/surfactant ratio can be used to explain the difference of
In addition, the researchers found that the optimum solubilization minimum-equilibrium IFT that may be observed with different
ratio follows a linear mixing rule: surfactant concentrations and WOR. This also explains why the

Opt vs Soap Fraction (theory) Opt vs Soap Fraction (exp) Opt vs Soap Fraction (theory) Opt vs Soap Fraction (exp)
10 10
Optiman Salinity

Optimal Salinity

1
1

0.1
0.1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
Xsoap Xsoap
((a)
a) (b)

Fig. 13—Relationship of optimal salinity and soap mole fraction by different acid-number methods for NI Blend and Yates oil
(Liu et al. 2010). (a) Nonaqueous phase titration; acid number = 0.75 mg KOH/g; (b) soap extraction by NaOH; acid number =
0.44 mg KOH/g.

898 December 2011 SPE Journal


0.4 phases (Hill et al. 1977; Pope et al. 1978; Glover et al. 1979; Gupta
Brine Oil Bank Transition Zone 1982). It was discovered that the surfactant micelles or microemul-
(Two Phases) sion droplets have an affinity for divalent ions similar to that of the
clays, and thus act as a flowing ion-exchange medium (Hirasaki

Two-Phases - II (+)

Two-Phases - II (−)
Three-Phases - III
1982; Hirasaki and Lawson 1986). The problem of divalent ions
0.3 is avoided by use of an alkali such as sodium carbonate or sodium
Residual Oil Saturation, PV

silicate (Holm and Robertson 1981).

Salinity Gradient. It was demonstrated that with a salinity


gradient:
0.2
1. Ahead of the active region, the system is overoptimum; sur-
factant is retarded by partitioning into the oil-phase.
2. The system passes through the active region of ultralow IFT
(Winsor III) where residual-oil displacement takes place.
3. Behind the active region, the system is underoptimum, with
Chemical Slug

lower-phase microemulsion, and the surfactant propagates with


0.1 the water velocity (Glover et al. 1979; Pope et al. 1979; Hirasaki
Polymer Drive et al. 1983). Thus, the salinity gradient tended to focus the sur-
factant near the advancing displacement front where salinity is
optimal and the phase behavior is Winsor III (Fig. 14). Also, the
salinity gradient helps to maintain polymer flow in the same phase
0 with the surfactant for the Winsor I conditions behind the active
0 0.4 0.8 1.2 1.6 region. The polymer is in the excess-brine phase in the Winsor II
Chemical Slug/Drive Injected, PV and III phase environments (Gupta 1982; Tham et al. 1983). The
example in Fig. 15 was injected overoptimum only for illustra-
Fig. 14—Oil saturation and surfactant production during ex- tion of surfactant transport with respect to salinity environment.
periment with finite, overoptimum surfactant slug and salinity Overoptimum salinity environments (Winsor II) can have viscous,
gradient (Hirasaki et al. 1983). high-internal-phase, water-in-oil emulsions [Hirasaki et al. (1983),
Fig. 14] that may be bypassed by the subsequent lower-salinity
fluids. In practice, the surfactant slug is injected in a near-opti-
mal to underoptimum salinity environment. Therefore, the gradi-
minimum IFT of small oil drops on a calcite plate occurred at the ent basically provides assurance that if overoptimum conditions
optimal salinity of the TC blend surfactant with zero soap fraction are unexpectedly reached during the process, the lower salinity
(i.e., 10–12% NaCl) [Zhang et al. (2006); Figs. 4 and 13]. The injected later will allow optimal conditions to be achieved and will
WOR of the small drops was very high. Equilibrium IFT measure- release surfactant trapped in the oil.
ments had lower optimal salinity because the soap/surfactant ratios
were larger owing to lower WORs. Soap/Surfactant Gradient. It was mentioned earlier that the
The dependence of optimal salinity on the soap/surfactant ratio optimal salinity changes as the soap/surfactant ratio changes.
also explains the transient minimum IFT observations in spinning- Thus, an alkaline/surfactant flood will have a gradient of optimal
drop measurement of a fresh oil drop in fresh surfactant solution salinity because of a gradient in the soap/surfactant ratio unless the
(Liu 2007). The soap concentration of the oil drop is changing soap content is negligible or the surfactant and soap have identical
as soap is being extracted from the small oil drop into the much optimal salinity (although surfactant would likely not be used if the
larger volume of surfactant solution. Thus, the soap/surfactant soap had a suitable optimal salinity). A gradient in the soap/sur-
concentration ratio of the oil drop changes from a large value to factant ratio exists because soap is generated in situ by interaction
near zero with time, and the minimum IFT (in time) occurs when between the alkali and the naphthenic acids in the crude oil, while
the soap/surfactant ratio of the oil drop corresponds to the ratio the synthetic surfactant is introduced with the injected fluid.
that is optimal for the salinity of the surfactant solution. The role of the soap/surfactant gradient in the ASP process was
evaluated with a 1D finite-difference simulator (Liu et al. 2008,
Composition Gradients 2010). Example composition and IFT profiles (Fig. 15) show the
Displacement of residual oil by surfactant flooding requires reduc- IFT dropping to ultralow values in a narrow region of the profile
ing the IFT to ultralow values such that disconnected oil droplets as the optimal salinity passes across the system salinity, which
can be mobilized. The ultralow IFT generally exists only in a nar- was constant in this example. There is only a short distance for
row salinity range near the optimal salinity. During the 1970s and the oil saturation to be reduced to a low value before the IFT
1980s, two schools of thought developed about how ultralow IFT again increases and traps any oil that has not been displaced. The
could be achieved in the displacement process (Gupta and Trush- oil saturation that will be trapped is approximately the saturation
enski 1979). One approach is to either preflush the formation to where the slope of the ultralow-IFT oil/water fractional-flow curve
reduce the formation salinity to a value near optimal, or to design becomes less than the dimensionless velocity of the displacement
the surfactant formulation such that the optimal salinity is equal to front (Pope 1980; Hirasaki 1981; Ramakrishnan and Wasan 1988,
the formation salinity, with the surfactant slug and drive injected 1989). Thus mobility control is important for displacement effi-
at the formation salinity (Maerker and Gale 1992). In the former ciency in addition to sweep efficiency for ASP flooding. Finite-
case, success was limited because the more viscous surfactant slug difference simulation showed that recovery decreased from 95 to
contacted portions of the reservoir that the preflush bypassed. In 86% as the aqueous viscosity decreased from 40 to 24 cp for oil
the latter case, this problem is avoided because there is no change with viscosity of 19 cp (Liu et al. 2008). This is consistent with a
in salinity because of dispersive mixing or crossflow. The other pair of experiments that differed only in polymer concentration.
approach is to have a salinity gradient such that the system has The effects of salinity, surfactant concentration, acid number,
overoptimum salinity ahead of, and underoptimum salinity behind, slug size, and dispersion on oil recovery are illustrated for a 0.2-
the active region. In this case, the salinity profile is certain to pass PV slug and laboratory-scale dispersion (Pe = 500) in Fig. 16 (Liu
through the optimal salinity somewhere in the displacement-front et al. 2010). The system is the one discussed in the IFT and phase-
region (Nelson 1981). behavior sections, and the black dot in Fig. 17 represents condi-
Whether the salinity is constant or a salinity gradient is used, tions of the successful sandpack experiment mentioned previously.
the electrolyte composition is further challenged by divalent ions in The effects of surfactant concentration and acid number (soap
the formation brine and ion exchange from the clays to the flowing content of the crude oil) are combined in a single parameter—the

December 2011 SPE Journal 899


Fig. 15—Profiles for large slug (0.5 PV) with low dispersion near optimal salinity (2% NaCl) (from Liu et al. 2010).

soap fraction at the waterflood residual-oil saturation. The range 2010). The lower salinity of the drive compensates for the lower
of salinity for greater-than-90% oil recovery is a function of the surfactant concentration such that the region of optimal salinity
soap fraction. The salinity for maximum oil recovery decreases again propagates with a near-unit velocity. In addition, injection
from optimal salinity of the surfactant to that of soap as the soap of the surfactant slug and polymer drive with a salinity that is less
fraction increases (straight line in the figure). The range of salini- that the optimal salinity of the surfactant alone makes it possible
ties for potentially high oil recovery is substantial, especially in to inject the surfactant slug with polymer without separation of the
the underoptimum region below the optimal line. surfactant and polymer into separate phases (Gupta 1982; Liu et al.
If the dispersion is increased to a representative field-scale 2008). Also, the salinity gradient avoids the large surfactant reten-
value [Pe = 50, Lake (1989)] with constant salinity and a 0.2-PV tion from microemulsion trapping by the polymer drive (Glover
slug, the region of greater-than-90% recovery all but disappears. et al. 1979; Hirasaki et al. 1983).
Dilution by mixing at the front and back of the surfactant slug low-
ers the surfactant concentration more than the soap concentration, Foam Mobility Control
and the propagation velocity of the soap/surfactant ratio for optimal Foam is usually considered as a means of mobility control for
salinity is greatly retarded. However, if the system is operated with gas-injection processes such as steam foam or CO2 foam. Foam
a salinity gradient, high oil recovery is again possible (Liu et al. mobility control for surfactant flooding is a natural progression

900 December 2011 SPE Journal


5.0

4.0
70% 50%
3.0 30%

Injected Salinity, %NaCl


2.0
90%

90%

1.0
Optimum
70%
Curve
50%
0.5

30%

0 0.2 0.4 0.6 0.8 1


Sor
X Soap

Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=4.0% NaCl (over-optimum)

Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=2.0% NaCl (near-optimum)

Acid No.=0.2mg/g, surfactant concentration=0.14%, salinity=1.0% NaCl (under -optimum)

Fig. 16—Recovery factor with small slug (0.2 PV) and low dispersion (Pe = 500) (Liu et al. 2010).

because the system already has surfactant present (Lawson and surfactant/polymer flooding (Kang et al. 2010). It has also been
Reisberg 1980). Moreover, at high temperatures, foam may be used as mobility control for surfactant aquifer remediation (Hira-
favored because polymer degradation is a concern (Srivastava and saki et al. 1997, 2000). Nonionic surfactants have been evaluated
Nguyen 2010). In fact, foam was used for mobility control for for mobility control of CO2 EOR (Adkins et al. 2010a, b). They
alkaline/surfactant flooding in China (Zhang et al. 2000; Wang can be injected dissolved in the CO2 phase and have less adsorp-
et al. 2001). Recently, it has been used to improve sweep in tion on carbonate formations compared with anionic surfactants.

Fig. 17—Displacement profiles for the displacement of MY residual crude oil by ASPF in 40-darcy sandpack (Li et al. 2010).

December 2011 SPE Journal 901


0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 2.0
Total PV
0 0.1 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
Liquid PV

NIP Air NIP Air NIP Air IOS Air IOS Air IOS Air IOS Air IOS Air IOS

1 ft/day 20 ft/day

Fig. 18—Profiles of the displacement of 266-cp Crude B with ASP and ASPF (Li et al. 2010).

If, in addition, the oil/water IFT can be reduced to ultralow values, Apparently, the viscous oil was being transported as an oil-in-water
a low-tension CO2 EOR process may be applicable for reservoirs emulsion with much less resistance than that of the crude oil.
with pressures below the minimum miscibility pressure.
Sweep of Layered Sands. Fig. 19 compares sweep in two cases
ASP Foam. The reduction of surfactant adsorption with alkali of a 19:1 permeability contrast layered sandpack initially filled
may result in the polymer being the most expensive chemical in with water dyed green. The sandpack is nearly completely swept
the ASP process. Experiments in 1D sandpacks have shown that with 1.0 TPV of surfactant alternating with gas (SAG) while the
an ASP process with the polymer drive replaced by a foam drive low-permeability layer is only one-quarter swept with water only
is equally efficient. Fig. 17 is an experiment in which the ASP (Li et al. 2010). The sweep efficiency is compared in Fig. 20 as
slug is alternated with equal-sized slugs of gas. The foam drive a function of the PV of liquid injected for SAG, water alternating
consists of slugs of the better-foaming surfactant component gas, and waterflooding.
(without polymer) alternated with equal-sized slugs of gas (Li
et al. 2010). Practically all of the 19-cp oil was recovered after 1.2 Potential for Fractured Formations. The improvement of sweep
TPV injected, but with only 0.6 PV of liquid injected. Experiments in layered sands suggests that foam may be helpful in the sweep
with different sands indicated that foam reduced mobility more of a system of fractures. Yan et al. (2006) showed that pregener-
in higher-permeability media, making it particularly attractive in ated foam does improve the sweep of parallel plates with different
layered systems. apertures to simulate heterogeneous, parallel fractures. Also, the
ASP foam was used to recover a 266-cp, 4.8-mg KOH/g TAN increased pressure gradient caused by foam flow in the fractures
crude oil (Fig. 18). What was remarkable is that the apparent increases the driving force for displacement of oil from the matrix
viscosity of the displacement process was only 80 cp or less. (Haugen et al. 2010; Abbasi-Asl et al. 2010). Farajzadeh et al.

SAG, 6 psi, fg=1/3 Water only, 4 psi

0.0 TPV

0.2 TPV

0.4 TPV

0.6 TPV

0.8 TPV

1.0 TPV

Fig. 19—Comparison of SAG with waterflood in 19:1 permeability ratio sandpack (Li et al. 2010).

902 December 2011 SPE Journal


SAG fg=2/3, 8psi
SAG fg=2/3, 6psi
SAG
SAG fg=4/5, 4psi
1.0
SAG fg=2/3, 4psi
WAG

Sweep Efficiency
0.8 SAG fg=3/4, 4psi
0.6 SAG fg=2/3, 2psi
Waterflood
SAG fg=1/3, 6psi
0.4
SAG fg=1/2, 4psi
0.2 WAG fg=4/5, 4psi
WAG fg=3/4, 4psi
0.0
0 0.5 1 1.5 2 2.5 3 WAG fg=2/3, 4psi
WAG fg=1/2, 4psi
PVs of Liquid Injected
Water fg=0, 4psi

Fig. 20—Sweep in 19:1 permeability contrast sandpack with SAG, WAG, and waterflood (Li et al. 2010).

(2010) describe how the viscous pressure gradient caused by foam 6. The soap generated in situ by the alkali causes a middle layer
flow in fractures can accelerate the production of oil that would to form and coexist with the lower-phase microemulsion, which
otherwise be produced from the matrix by gravity drainage. results in ultralow IFT over a wide range of salinity.
7. Anionic surfactants and sodium carbonate can alter wettability
Field Pilots for either sandstone or carbonate formations. Spontaneous oil
A pilot of the alkaline/surfactant process is described by Falls et al. displacement can occur by gravity drainage.
(1994). This pilot was tested without polymer, with the intention 8. Foam can be used as the drive of the alkaline/surfactant process
of a subsequent test with polymer. Nevertheless, the interpretation in place of the polymer drive.
of induction logs suggested 100% displacement efficiency in the 9. Foam can efficiently sweep layered and fractured systems.
region swept by the injected fluids.
A refinement of the alkaline/surfactant process with recent Acknowledgments
understanding about the IOS is described by Buijse et al. (2010). The authors acknowledge the financial support by DOE grant
The paper also discusses the importance of the crude-oil composi- DE-FC26-03NT15406 and the Rice University Consortium on
tion. The process was tested in the field with a single well chemi- Processes in Porous Media. The information and insight we gained
cal-tracer test. Stoll et al. (2010) describe pilot tests in Oman. from our long collaboration with Gary Pope and Kishore Mohanty
Gao and Gao (2010) summarized the pilots in Daqing oil are also acknowledged.
field.
References
Conclusions Abbasi-Asl, Y., Pope, G.A., and Delshad, M. 2010. Mechanistic Modeling
The technology of surfactant flooding has advanced to overcome of Chemical Transport in Naturally Fractured Oil Reservoirs. Paper
many of the past causes of failures and to reduce the amount of sur- SPE 129661 presented at the SPE Improved Oil Recovery Symposium,
factant required. These developments are summarized as follows: Tulsa, 24–28 April. doi: 10.2118/129661-MS.
1. Surfactant adsorption can be significantly reduced in sandstone Abe, M., Schechter, D., Schechter, R.S., Wade, W.H., Weerasooriya, U.,
and carbonate formations by injection of an alkali such as and Yiv, S. 1986. Microemulsion Formation with Branched Tail Poly-
sodium carbonate. The alkali also sequesters divalent ions. The oxyethelene Sulfonate Surfactants. Journal of Colloid and Interface
reduced adsorption permits lower surfactant concentrations. Science 114 (2): 342–356. doi: 10.1016/0021-9797(86)90420-0.
2. A wide selection of surfactant structures is now available to meet Adams, W.T. and Schievelbein, V.H. 1987. Surfactant Flooding Carbon-
requirements for specific applications. ate Reservoirs. SPE Res Eng 2 (4): 619–626. SPE-12686-PA. doi:
a. Branched alcohol alkoxylate sulfates and sulfonates are toler- 10.2118/12686-PA.
ant of divalent ions. Ethoxylation increases optimal salinity; Adibhatia, B. and Mohanty, K.K. 2007. Simulation of Surfactant-Aided
propoxylation decreases optimal salinity. In both cases, Gravity Drainage in Fractured Carbonates. Paper SPE 106161 pre-
EO or PO, the optimal salinity decreases with increasing sented at the SPE Reservoir Simulation Symposium, Houston, 26–28
temperature. February. doi: 10.2118/106161-MS.
b. Alkyloxylated glycidyl ether sulfonate is more expensive Adibhatla, B. and Mohanty, K.K. 2008. Oil Recovery From Fractured
than sulfate but is stable at elevated temperatures. Carbonates by Surfactant-Aided Gravity Drainage: Laboratory Experi-
c. IOSs are low-cost, double-tailed surfactants. ments and Mechanistic Simulations. SPE Res Eval & Eng 11 (1):
3. Aqueous solutions of a blend of N67-7PO sulfate and IOS1518 119–130. SPE-99773-PA. doi: 10.2118/99773-PA.
with alkali have a larger single-phase region extending to higher Adkins, S.S., Chen, X., Chan, I., Torino, E., Nguyen, Q.P., Sanders, A.,
salinities and calcium-ion concentrations than either alone. and Johnston, K.P. 2010a. Morphology and Stability of CO2-in-Water
This blend, without alcohol, can form a single phase for injec- Foams with Nonionic Hydrocarbon Surfactants. Langmuir 26 (8):
tion with polymer but can form microemulsions with crude oil 5335–5348. doi: 10.1021/la903663v.
without forming a gel. Adkins, S.S., Chen, X., Nguyen, Q.P., Sanders, A.W., and Johnston, K.P.
4. Soap generated in situ by the alkali is a cosurfactant that can 2010b. Effect of branching on the interfacial properties of nonionic
change the phase behavior of the injected surfactant solution hydrocarbon surfactants at the air-water and carbon dioxide-water
from lower- to middle- to upper-phase microemulsions. It is interfaces. Journal of Colloid and Interface Science 346 (2): 455–463.
lower phase when injected, middle phase at the displacement doi: 10.1016/j.jcis.2009.12.059.
front, and upper phase ahead of the displacement front. Anderson, G.A., Delshad, M., King, C.B., Mohammadi, H., and Pope, G.A.
5. Injection of the surfactant and polymer at salinity that is under- 2006. Optimization of Chemical Flooding in a Mixed-Wet Dolomite
optimum with respect to the injected surfactant avoids surfac- Reservoir. Paper SPE 100082 presented at the SPE/DOE Symposium on
tant/polymer phase separation and microemulsion trapping. Improved Oil Recovery, Tulsa, 22–26 April. doi: 10.2118/100082-MS.

December 2011 SPE Journal 903


Anderson, W.G. 1987a. Wettability Literature Survey— Part 4: Effects of Delshad, M., Najafabadi, N.F., Anderson, G.A., Pope, G.A., and Sepehr-
Wettability on Capillary Pressure. J Pet Technol 39 (10): 1283–1300. noori, K. 2006. Modeling Wettability Alteration in Naturally Fractured
SPE-15271-PA. doi: 10.2118/15271-PA. Reservoirs. Paper SPE 100081 presented at the SPE/DOE Symposium
Anderson, W.G. 1987b. Wettability Literature Survey –Part 5: The Effects on Improved Oil Recovery, Tulsa, 22–26 April. doi: 10.2118/100081-
of Wettability on Relative Permeability. J Pet Technol 39 (11): 1453– MS.
1468. SPE-16323-PA. doi: 10.2118/16323-PA. Dubey, S.T. and Doe, P.H. 1993. Base Number and Wetting Properties
Atkinson, H. 1927. Recovery of Petroleum from Oil Bearing Sands. US of Crude Oil. SPE Res Eng 8 (3): 195–200. SPE-22598-PA. doi:
Patent No. 1,651,311. 10.2118/22598-PA.
Austad, T., Matre, B., Milter, J., Sævareid, A., and Øyno, L. 1998. Chemi- Dwarakanath, V., Chaturvedi, T., Jackson, A.C., Malik, T., Siregar, A., and
cal flooding of oil reservoirs 8. Spontaneous oil expulsion from oil- and Zhao, P. 2008. Using Co-Solvents to Provide Gradients and Improve
water-wet low permeable chalk material by imbibition of aqueous sur- Oil Recovery During Chemical Flooding in a Light Oil Reservoir. Paper
factant solutions. Colloids and Surfaces A: Physico. Eng. Aspects 137 SPE 113965 presented at the SPE/DOE Symposium on Improved Oil
(1–3): 117–129. doi: 10.1016/S0927-7757(97)00378-6. Recovery, Tulsa, 19–23 April. doi: 10.2118/113965-MS.
Babadagli, T. 2001. Scaling of Co-current and Counter-Current Capillary Ehrlich, R. and Wygal, R.J. 1977. Interrelation of Crude Oil and Rock
Imbibition for Surfactant and Polymer Injection in Naturally Fractured Res- Properties with the Recovery of Oil by Caustic Waterflooding. SPE J.
ervoirs. SPE J. 6 (4): 465–478. SPE-74702-PA. doi: 10.2118/74702-PA. 17 (4): 263–270. SPE-5830-PA. doi: 10.2118/5830-PA.
Bae, J.H. and Petrick, C.B. 1977. Adsorption/Retention of Petroleum Ehrlich, R., Hasiba, H.H., and Raimondi, P. 1974. Alkaline Waterflooding
Sulfonate in Berea Cores. SPE J. 17 (5): 353–357. SPE-5819-PA. doi: for Wettability Alteration-Evaluating a Potential Field Application.
10.2118/5819-PA. J Pet Technol 26 (12): 1335–1342. SPE-4905-PA. doi: 10.2118/
Barnes, J.R., Dirkzwager, H., Smit, J.R., Smit, J.P., On, A., Navarrete, R.C., 4905-PA.
Ellison, B.H., and Buijse, M.A. 2010. Application of Internal Olefin Falls, A.H., Thigpen, D.R., Nelson, R.C., Ciaston, J.W., Lawson, J.B.,
Sulfonates and Other Surfactants to EOR. Part 1: Structure—Perfor- Good, P.A., Ueber, R.C., and Shahin, G.T.1994. Field Test of Cosur-
mance Relationships for Selection at Different Reservoir Conditions. factant-Enhanced Alkaline Flooding. SPE Res Eng 9 (3): 217–223.
Paper SPE 129766 presented at the SPE Improved Oil Recovery Sym- SPE-24117-PA. doi: 10.2118/24117-PA.
posium, Tulsa, 24–28 April. doi: 10.2118/129766-MS. Fan, T. and Buckley, J.S. 2007. Acid Number Measurements Revisited. SPE
Barnes, J.R., Smit, J.P., Smit, J.R., Shpakoff, P.G., Raney, K.H., and J. 12 (4): 496–500. SPE-99884-PA. doi: 10.2118/99884-PA.
Puerto, M.C. 2008. Development of Surfactants for Chemical Flooding Farajzadeh, R., Wassing, B., and Boerrigter, P. 2010. Foam Assisted Gas Oil
at Difficult Reservoir Conditions. Paper SPE 113313 presented at the Gravity Drainage in Naturally-Fractured Reservoirs. Paper SPE 134203
SPE/DOE Symposium on Improved Oil Recovery, Tulsa, 20–23 April. presented at the SPE Annual Technical Conference and Exhibition,
doi: 10.2118/113313-MS. Florence, Italy, 19–27 September. doi: 10.2118/134203-MS.
Baviere, M., Glenat, P., Plazanet, V., and Labrid, J. 1995. Improved EOR Flaaten, A.K., Nguyen, Q.P., Pope, G.A., and Zhang, J. 2009. A System-
by Use of Chemicals in Combination. SPE Res Eng 10 (3): 187–193. atic laboratory Approach to Low-Cost, High-Performance Chemical
SPE-27821-PA. doi: 10.2118/27821-PA. Flooding. SPE Res Eval & Eng 12 (5): 713–723. SPE-113469-PA. doi:
Benton, W.J. and Miller, C.A. 1983. Lyotropic liquid crystalline phases 10.2118/113469-PA.
and dispersions in dilute anionic surfactant-alcohol-brine systems. 1. Foster, W.R. 1973. A Low Tension Waterflooding Process. J Pet Technol
Patterns of phase behavior. J. Phys. Chem. 87 (24): 4981–4991. doi: 25 (2): 205–210; Trans., AIME, 255. SPE-3803-PA. doi: 10.2118/3803-
10.1021/j150642a042. PA.
Bourrel, M. and Schechter, R.S. 1988. Microemulsions and Related Sys- Gao, S. and Gao, Q. 2010. Recent Progress and Evaluation of ASP Flood-
tems: Formulation, Solvency, and Physical Properties, Vol. 30. New ing for EOR in Daqing Oil Field. Paper SPE 127714 presented at the
York: Surfactant Science Series, Marcel Dekker. SPE EOR Conference at Oil & Gas West Asia, Muscat, Oman, 11–13
Bryan, J, Mai, A., and Kantzas, A. 2008. Investigation into the Process April. doi: 10.2118/127714-MS.
Responsible for Heavy Oil Recovery by Alkali-Surfactant Flooding. Ghosh, O. 1985. Liquid Crystal to Microemulsion Transitions in Anionic
Paper SPE 113993 presented at the SPE/DOE Symposium on Improved Surfactant-Oil-Brine Systems. PhD thesis, Rice University, Ann Arbor,
Oil Recovery, Tulsa, 20–23 April. doi: 10.2118/113993-MS. Michigan (October 1985).
Buckley, J.S., Takamura, K., and Morrow, N.R. 1989. Influence of Electric Glover, C.J., Puerto, M.C., Maerker, J.M., and Sandvik, E.L. 1979. Sur-
Surface Charges on the Wetting Properties of Crude Oils. SPE Res Eng factant Phase Behavior and Retention in Porous Media. SPE J. 19 (3):
4 (3): 332–340. SPE-16964-PA. doi: 10.2118/16964-PA. 183–193. SPE-7053-PA. doi: 10.2118/7053-PA.
Buijse, M.A., Prelicz, R.M., Barnes, J.R., and Cosmo, C. 2010. Applica- Gogarty, W.B. 1977. Oil recovery with surfactants: History and a current
tion of Internal Olefin Sulfonates and Other Surfactants to EOR. Part appraisal. In Improved Oil Recovery by Surfactant and Polymer Flood-
2: The Design and Execution of an ASP Field Test. Paper SPE 129769 ing, ed. D.O. Shah and R.S. Schechter, 27–54. New York: Academic
presented at the SPE Improved Oil Recovery Symposium, Tulsa, 24–28 Press.
April. doi: 10.2118/129769-MS. Gupta, R., Mohan, K., and Mohanty, K.K. 2009. Surfactant Screening for
Cayias, J.L., Schechter, R.S., and Wade, W.H. 1977. The utilization of Wettability Alteration in Oil-Wet Fractured Carbonates. Paper SPE
petroleum sulfonates for producing low interfacial tensions between 124822 presented at the SPE Annual Technical Conference and Exhibi-
hydrocarbons and water. J. Colloid Interface Sci. 59 (1): 31–38. doi: tion, New Orleans, 19–27 September. doi: 10.2118/124822-MS.
10.1016/0021-9797(77)90335-6. Gupta, R. and Mohanty, K. 2010. Temperature Effects on Surfactant-Aided
Chen, H.L., Lucas, L.R., Nogaret, L.A.D, Yang, H.D., and Kenyon, D.E. Imbibition Into Fractured Carbonates. SPE J. 15 (3): 588-597. SPE-
2000. Laboratory Monitoring of Surfactant Imbibition Using Com- 110204-PA. doi: 10.2118/110204-PA.
puterized Tomography. Paper SPE 59006 presented at the SPE/DOE Gupta, R. and Mohanty, K.K. 2008. Wettability Alteration of Fractured
International Petroleum Conference and Exhibition in Mexico, Villa- Carbonate Reservoirs. Paper SPE 113407 presented at SPE/DOE
hermosa, Mexico, 1–3 February. doi: 10.2118/59006-MS. Symposium on Improved Oil Recovery, Tulsa, 20–23 April. doi:
Cheng, K.H. 1986. Chemical Consumption During Alkaline Flood- 10.2118/113407-MS.
ing: A Comparative Evaluation. Paper SPE 14944 presented at the Gupta, S.P. 1982. Dispersive Mixing Effects on the Sloss Field Micellar
SPE Enhanced Oil Recovery Symposium, Tulsa, 20–23 April. doi: System. SPE J. 22 (4): 481–492. SPE-9782-PA. doi: 10.2118/9782-
10.2118/14944-MS. PA.
Cuiec, L. 1977. Study of Problems Related to the Restoration of the Natural Gupta, S.P. and Trushenski, S.P. 1979. Micellar Flooding—Compositional
State of Core Samples. J Can Pet Technol 16 (4): 68–80. JCPT Paper Effects on Oil Displacement. SPE J. 19 (2): 116–128; Trans., AIME,
No. 77-04-09. doi: 10.2118/77-04-09. 267. SPE-7063-PA. doi: 10.2118/7063-PA.
Delshad, M., Najafabadi, N.F., and Sepehnoori, K. 2009. Scale Up Method- Haugen, A., Ferno, M.A., Graue, A., and Bertin, H.J. 2010. Experimental
ology for Wettability Modification in Fractured Carbonates. Paper SPE Study of Foam Flow in Fractured Oil-Wet Limestone for Enhanced
118915 presented at the SPE Reservoir Simulation Symposium, The Oil Recovery. Paper SPE 129763 presented at the SPE Improved Oil
Woodlands, Texas, 2–4 February. doi: 10.2118/118915-MS. Recovery Symposium, Tulsa, 24–28 April. doi: 10.2118/129763-MS.

904 December 2011 SPE Journal


Healey, R.N. and Reed, R.L. 1974. Physicochemical Aspects of Micro- Kang, W., Liu, S., Meng, L., Cao, D., and Fan, H. 2010. A Novel Ultra-low
emulsion Flooding. SPE J. 14 (5): 491–501; Trans., AIME, 265. SPE- Interfacial Tension Foam Flooding Agent to Enhance Heavy Oil Recov-
4583-PA. doi: 10.2118/4583-PA. ery. Paper SPE 129175 presented at the SPE Improved Oil Recovery
Healy, R.N., Reed, R.L., and Stenmark, D.K. 1976. Multiphase Microemul- Symposium, Tulsa, 24–28 April. doi: 10.2118/129175-MS.
sion Systems. SPE J. 16 (3): 147–160; Trans., AIME, 261. SPE-5565- Kumar, R. and Mohanty, K.K. 2010. ASP Flooding of Viscous Oil. Paper
PA. doi: 10.2118/5565-PA. SPE 135265 presented at the SPE Annual Technical Conference and
Hill, H.J., Helfferich, F.G., Lake, L.W., and Reisberg, J. 1977. Cation Exhibition, Florence, Italy, 19–27 September. doi: 10.2118/135265-
Exchange and Chemical Flooding. J Pet Technol 29 (10): 1336–1338. MS.
SPE-6642-PA. doi: 10.2118/6642-PA. Kumar, R., Dao, E., and Mohanty, K.K. 2010. Emulsion Flooding of Heavy
Hill, H.J., Reisberg, J., and Stegemeier, G.L. 1973. Aqueous Surfactant Oil. Paper SPE 129914 presented at the SPE/DOE Enhanced Oil Recov-
Systems For Oil Recovery. J Pet Techol 25 (2): 186–194. SPE-3798- ery Symposium, Tulsa, 20–23 April. doi: 10.2118/129914-MS.
PA. doi: 10.2118/3798-PA. Lake, L.W. 1989. Enhanced Oil Recovery, 166. Englewood Cliffs, New
Hirasaki, G. and Zhang, D.L. 2004. Surface Chemistry of Oil Recovery Jersey: Prentice Hall.
From Fractured, Oil-Wet, Carbonate Formation. SPE J. 9 (2): 151–162. Lawson, J.B. and Reisberg, J. 1980. Alternate Slugs of Gas and Dilute
SPE-88365-PA. doi: 10.2118/88365-PA. Surfactant for Mobility Control During Chemical Flooding. Paper SPE
Hirasaki, G., Miller C.A., Pope, G.A., and Jackson, R.E. 2005. Surfactant 8839 presented at the SPE/DOE Enhanced Oil Recovery Symposium,
Based Enhanced Oil Recovery and Foam Mobility Control. 2nd Annual Tulsa, 20–23 April. doi: 10.2118/8839-MS.
Technical Report (July 2004–June 2005), Contract No. DE-FC26- Levitt, D.B., Jackson, A.C., Heinson, C., Britton, L.N., Malik, T., Dwara-
03NT15406, US DOE, Washington, DC (July 2005). kanath, V., and Pope, G.A. 2009. The Identification and Evaluation
Hirasaki, G.J. 1981. Application of the Theory of Multicomponent, Multi- of High-Performance EOR Surfactants. SPE Res Eval & Eng 12 (2):
phase Displacement to Three-Component, Two-Phase Surfactant Flood- 243–253. SPE-100089-PA. doi: 10.2118/100089-PA.
ing. SPE J. 21 (2): 191–204. SPE-8373-PA. doi: 10.2118/8373-PA. Li, R.F., Yan, W., Liu, S., Hirasaki, G., and Miller, C.A. 2010. Foam
Hirasaki, G.J. 1982. Ion Exchange With Clays in the Presence of Surfactant. Mobility Control for Surfactant Enhanced Oil Recovery. SPE J. 15 (4):
SPE J. 22 (2): 181–192. SPE-9279-PA. doi: 10.2118/9279-PA. 928-942. SPE-113910-PA. doi: 10.2118/113910-PA.
Hirasaki, G.J. and Lawson, J.B. 1986. An Electrostatic Approach to the Liu, Q., Dong, M., and Ma, S. 2006. Alkaline/Surfactant Flood Potential in
Association of Sodium and Calcium with Surfactant Micelles. SPE Western Canadian Heavy Oil Reservoirs. Paper SPE 99791 presented
Res Eng 1 (2): 119–130; Trans., AIME, 281. SPE-10921-PA. doi: at the SPE/DOE Symposium on Improved Oil Recovery, Tulsa, 22–26
10.2118/10921-PA. April. doi: 10.2118/99791-MS.
Hirasaki, G. J. et al. 1997. Field Demonstration of the Surfactant/Foam Liu, S. 2007. Alkaline Surfactant Polymer Enhanced Oil Recovery Pro-
Process for Aquifer Remediation. SPE 39292 presented at the SPE cesses. Ph.D. thesis, Rice University, Ann Arbor, Michigan (Decem-
Annual Technical Conference and Exhibition, San Antonio, 5–8 Octo- ber 2007). http://www.owlnet.rice.edu/~gjh/Consortium/Manuscripts/
ber. doi: 10.2118/39292-MS. Liu_Thesis.pdf.
Hirasaki, G.J., Jackson, R.E., Jin, M., Lawson, J.B., Londergan, J., Liu, S., Feng Li, R., Miller, C.A., and Hirasaki, G.J. 2010. Alkaline/Surfac-
Meinardus, H., Miller, C.A., Pope, G.A., Szafranski, R., and Tanzil, tant/Polymer Processes: Wide Range of Conditions for Good Recovery.
D. 2000. Field Demonstration of the Surfactant/Foam Process for SPE J. 15 (2): 282-293. SPE-113936-PA. doi: 10.2118/113936-PA.
Remediation of a Heterogeneous Aquifer Contaminated with DNAPL. Liu, S., Zhang, D.L., Yan, W., Puerto, M., Hirasaki, G.J., and Miller, C.A.
In NAPL Removal: Surfactants, Foams, and Microemulsions, ed. S. 2008. Favorable Attributes of Alkali-Surfactant-Polymer Flooding. SPE
Fiorenza, C.A. Miller, C.L. Oubre, and C.H. Ward, Part 1, 3–163. J. 13 (1): 5–16. SPE-99744-PA. doi: 10.2118/99744-PA.
AATDF monograph series, Boca Raton, Florida, USA: Lewis Publish- Lopez-Salinas, J.L., Miller, C.A., Koo, K.H.K., Puerto, M., and Hirasaki,
ers, CRC Press. G.J. 2009 Viscometer for opaque, sealed microemulsion samples.
Hirasaki, G.J., van Domselaar, H.R., and Nelson, R.C. 1983. Evaluation of Paper SPE 121575 presented at the SPE International Symposium on
the Salinity Gradient Concept in Surfactant Flooding. SPE J. 23 (3): Oilfield Chemistry, The Woodlands, Texas, USA, 20–22 April. doi:
486–500. SPE-8825-PA. doi: 10.2118/8825-PA. 10.2118/121575-MS.
Høgnesen, E.J., Olsen, M., and Austad, T. 2006. Capillary and Gravity Lyklema, J. 1995. Fundamentals of Interface and Colloid Science. Volume
Dominated Flow Regimes in Displacement of Oil from an Oil-Wet II: Solid-Liquid Interfaces, Appendix 3, A3.1–A3.6. San Diego, Cali-
Chalk Using Cationic Surfactant. Energy Fuels 20 (3): 1118–1122. fornia: Academic Press.
doi: 10.1021/ef050297s. Maerker, J.M. and Gale, W.W. 1992. Surfactant Flood Process Design for
Høgnesen, E.J., Standnes, D.C., and Austad, T. 2004. Scaling Sponta- Loudon. SPE Res Eng 7 (1): 36–44; Trans., AIME, 293. SPE-20218-
neous Imbibition of Aqueous Surfactant Solution Into Preferential PA. doi: 10.2118/20218-PA.
Oil-Wet Carbonates. Energy Fuels 18 (6): 1665–1675. doi: 10.1021/ Melrose, J.C. and Brandner, C.F. 1974. Role of Capillary Forces In
ef040035a. Detennining Microscopic Displacement Efficiency For Oil Recovery
Holm, L.W. 1977. Soluble oils for improved oil recovery. In Improved Oil By Waterflooding. J Can Pet Technol 13 (4): 54–62. JCPT Paper No.
Recovery by Surfactant and Polymer Flooding, ed. D.O. Shah and R.S. 74-04-05. doi: 10.2118/74-04-05.
Schechter, 453–485. New York: Academic Press. Miller, C.A., Ghosh, O., and Benton, W.J. 1986. Behavior of dilute lamellar
Holm, L.W. and Robertson, S.D. 1981. Improved Micellar/Polymer Flood- liquid-crystalline phases. Colloids and Surfaces 19 (2–3): 197–223. doi:
ing With High-pH Chemicals. J Pet Technol 33 (1): 161–171. SPE- 10.1016/0166-6622(86)80336-5.
7583-PA. doi: 10.2118/7583-PA. Milter, J. and Austad, T. 1996a. Chemical flooding of oil reservoirs
Huh, C. 1979. Interfacial tensions and solubilizing ability of a micro- 6. Evaluation of the mechanism for oil expulsion by spontaneous
emulsion phase that coexists with oil and brine. Journal of Col- imbibition of brine with and without surfactant in water-wet, low-
loid and Interface Science 71 (2): 408–426. doi: 10.1016/0021- permeable, chalk material. Colloids and Surfaces A: Physicochemi-
9797(79)90249-2. cal and Engineering Aspects 113 (3): 269–278. doi: 10.1016/0927-
Jadhunandan, P.P. and Morrow, N.R. 1995. Effect of Wettability on Water- 7757(96)03631-X.
flood Recovery for Crude-Oil/Brine/Rock Systems. SPE Form Eval 10 Milter, J. and Austad, T. 1996b. Chemical flooding of oil reservoirs 7.
(1): 40–46. SPE-22597-PA. doi: 10.2118/22597-PA. Oil expulsion by spontaneous imbibition of brine with and without
Jennings, H.Y. Jr. 1975. A Study of Caustic Solution-Crude Oil Interfacial surfactant in mixed-wet, low permeability chalk material. Colloids
Tensions. SPE J. 15 (3): 197–202. SPE-5049-PA. doi: 10.2118/5049- and Surfaces A: Physicochemical and Engineering Aspects 117 (1–2):
PA. 109–115. doi: 10.1016/0927-7757(96)03693-X.
Jensen, J.A. and Radke, C.J. 1988. Chromatographic Transport of Alkaline Mohammadi, H., Delshad, M., and Pope, G.A. 2009. Mechanistic Model-
Buffers Through Reservoir Rock. SPE Res Eng 3 (3): 849–856. SPE- ing of Alkaline/Surfactant/Polymer Floods. SPE Res Eval & Eng 12
14295-PA. doi: 10.2118/14295-PA. (4): 518–527. SPE-110212-PA. doi: 10.2118/110212-PA.
Johnson, C.E. Jr. 1976. Status of Caustic and Emulsion Methods. J Pet Morrow, N.R. 1990. Wettability and Its Effect on Oil Recovery. J Pet Tech-
Technol 28 (1): 85–92. SPE-5561-PA. doi: 10.2118/5561-PA. nol 24 (12): 1476–1484. SPE-21621-PA. doi: 10.2118/21621-PA.

December 2011 SPE Journal 905


Morrow, N.R. and Mason, G. 2001. Recovery of Oil by Spontaneous Imbi- Seethepalli, A. , Adibhatla, B., and Mohanty, K.K. 2004. Physicochemical
bition. Current Opinion in Colloid & Interface Science 6 (4): 321–337. Interactions During Surfactant Flooding of Fractured Carbonate Res-
doi: 10.1016/S1359-0294(01)00100-5. ervoirs. SPE J. 9 (4): 411–418. SPE-89423-PA. doi: 10.2118/89423-
Najafabadi, N.F., Delshad, M., Sepehrnoori, K., Nguyen, Q.P., and Zhang, PA.
J. 2008. Chemical Flooding of Fractured Carbonates Using Wettability Somasundaran, P. and Zhang, L. 2006. Adsorption of surfactant on minerals
Modifiers. Paper SPE 113369 presented at the SPE/DOE Symposium for wettability control in improved oil recovery processes. J Pet Sci Eng
on Improved Oil Recovery, Tulsa, 20–23 April. doi: 10.2118/113369- 52 (1–4): 198–212. doi: 10.1016/j.petrol.2006.03.022.
MS. Southwick, J.G. 1985. Solubility of Silica in Alkaline Solutions: Implica-
Nelson, R.C. 1981. Further Studies on Phase Relations in Chemical Flood- tions for Alkaline Flooding, SPE J. 25 (6): 857–864. SPE-12771-PA.
ing. In Surface Phenomena in Enhanced Oil Recovery, ed. D.O. Shah, doi: 10.2118/12771-PA.
73–104. New York: Plenum Publishing. Srivastava, M. and Nguyen, Q.P. 2010. Application of Gas for Mobility
Nelson, R.C. and Pope, G.A. 1978. Phase Relationships in Chemical Control in Chemical EOR in Problematic Carbonate Reservoirs. Paper
Flooding. SPE J. 18 (5): 325–338; Trans., AIME, 265. SPE-6773-PA. SPE 129840 presented at the SPE Improved Oil Recovery Symposium,
doi: 10.2118/6773-PA. Tulsa, 24–28 April. doi: 10.2118/129840-MS.
Nelson, R.C., Lawson, J.B., Thigpen, D.R., and Stegemeier, G.L. 1984. Standnes, D.C. and Austad, T. 2000. Wettability alteration in chalk 2.
Cosurfactant-Enhanced Alkaline Flooding. Paper SPE 12672 presented Mechanism for wettability alteration from oil-wet to water-wet using
at the SPE Enhanced Oil Recovery Symposium, Tulsa, 15–18 April. surfactants. J. Pet. Sci. Eng. 28 (3): 123–143. doi: 10.1016/S0920-
doi: 10.2118/12672-MS. 4105(00)00084-X.
Novosad, Z. and Novosad, J. 1984. Determination of Alkalinity Losses Standnes, D.C., Nogaret, L.A.D., Chen, H.-C., and Austad, T. 2002. An
Resulting From Hydrogen Ion Exchange in Alkaline Flooding. SPE J. Evaluation of Spontaneous Imbibition of Water into Oil-Wet Carbonate
24 (1): 49–52. SPE-10605-PA. doi: 10.2118/10605-PA. Reservoir Cores Using a Nonionic and a Cationic Surfactant. Energy
Pennell, K.D., Pope, G.A., and Abriola, L.M. 1996. Influence of Viscous Fuels 16 (6): 1557–1564. doi: 10.1021/ef0201127.
and Buoyancy Forces on the Mobilization of Residual Tetrachloro- Stegemeier, G.L. 1977. Mechanisms of entrapment and mobilization of
ethylene during Surfactant Flushing. Environ. Sci. Technol. 30 (4): oil in porous media. In Improved Oil Recovery by Surfactant and
1328–1335. doi: 10.1021/es9505311. Polymer Flooding, ed. D.O. Shah and R.S. Schechter, 55–91. New
Peru, D.A. and Lorenz, P.B. 1990. Surfactant-Enhanced Low-pH Alkaline York: Academic Press.
Flooding. SPE Res Eng 5 (3): 327–332. doi: 10.2118/17117-PA. Stoll, W.M., al Shureqi, H., Finol, J., Al-Harthy, A.A., Oyemade, S., de
Pope, G.A. 1980. The Application of Fractional Flow Theory to Kruijf, A., van Wunnik, J., Arkesteijn, F., Bouwmeester, R., and Faber,
Enhanced Oil Recovery. SPE J. 20 (3): 191–205. SPE-7660-PA. doi: M.J. 2010. Alkaline-Surfactant-Polymer Flood: From the Laboratory
10.2118/7660-PA. to the Field. Paper SPE 129164 presented at the SPE EOR Confer-
Pope, G.A., Lake, L.W., and Helfferich, F.G. 1978. Cation Exchange in ence at Oil & Gas West Asia, Muscat, Oman, 11–13 April. doi:
Chemical Flooding: Part 1—Basic Theory Without Dispersion. SPE J. 10.2118/129164-MS.
18 (6): 418–434. SPE-6771-PA. doi: 10.2118/6771-PA. Surkalo, H. 1990. Enhanced Alkaline Flooding. J Pet Technol 42 (1): 6–7.
Pope, G.A., Wang, B., and Tsaur, K. 1979. A Sensitivity Study of Micel- SPE-19896-PA. doi: 10.2118/19896-PA.
lar/Polymer Flooding. SPE J. 19 (6): 357–368. SPE-7079-PA. doi: Sydansk, R.D. 1982. Elevated-Temperature Caustic/Sandstone Interaction:
10.2118/7079-PA. Implications for Improving Oil Recovery. SPE J. 22 (4): 453–462. SPE-
Puerto, M., Hirasaki, G.J., and Miller, C.A. 2010. Surfactant Systems for 9810-PA. doi: 10.2118/9810-PA.
EOR in High-Temperature, High-Salinity Environments. Paper SPE Tabatabal, A., Gonzalez, M.V., Harwell, J.H., and Scamehorn, J.F. 1993.
129675 presented at SPE Improved Oil Recovery Symposium, Tulsa, Reducing Surfactant Adsorption in Carbonate Reservoirs. SPE Res Eng
24–28 April. doi: 10.2118/129675-MS. 8 (2): 117–122. SPE-24105-PA. doi: 10.2118/24105-PA.
Qutubuddin, S., Miller, C.A., Benton, W.J., and Fort, T. Jr. 1985. Effects Taber, J.J. 1969. Dynamic and Static Forces Required To Remove a Discon-
of polymers, electrolytes, and pH on microemulsion phase behavior. In tinuous Oil Phase from Porous Media Containing Both Oil and Water.
Macro- and Microemulsions: Theory and Applications, ed. D.O. Shah, SPE J. 9 (1): 3–12. SPE-2098-PA. doi: 10.2118/2098-PA.
No. 272, 223–251. Washington, DC: Symposium Series, ACS. Takamura, K. and Chow, R.S. 1985. The electric properties of the bitu-
Ramakrishnan, T.S. and Wasan, D.T. 1988. The Role of Adsorption in men/water interface Part II. Application of the ionizable surface-
High pH Flooding. Paper SPE 17408 presented at the SPE California group model. Colloids and Surfaces 15: 35–48. doi: 10.1016/0166-
Regional Meeting, Long Beach, California, USA, 23–25 March. doi: 6622(85)80053-6.
10.2118/17408-MS. Talley, L.D. 1988. Hydrolytic Stability of Alkylethoxy Sulfates. SPE Res
Ramakrishnan, T.S. and Wasan, D.T. 1989. Fractional-Flow Model for Eng 3 (1): 235–242. SPE-14912-PA. doi: 10.2118/14912-PA.
High-pH Flooding. SPE Res Eng 4 (1): 59–68. SPE-14950-PA. doi: Tham, M.J., Nelson, R.C., and Hirasaki, G.J. 1983. Study of the Oil Wedge
10.2118/14950-PA. Phenomenon Through the Use of a Chemical Flood Simulator. SPE J.
Reed, R.L. and Healy, R.N. 1977. Some physico-chemical aspects of 23 (5): 746–758. SPE-10729-PA. doi: 10.2118/10729-PA.
microemulsion flooding: a review. In Improved Oil Recovery by Treiber, L.E. and Owens, W.W. 1972. A Laboratory Evaluation of the
Surfactant and Polymer Flooding, ed. D.O. Shah and R.S. Schechter, Wettability of Fifty Oil-Producing Reservoirs. SPE J. 12 (6): 531–540.
383–437. New York: Academic Press. SPE-3526-PA. doi: 10.2118/3526-PA.
Sahni, V., Dean, R.M., Britton, C., Weerasooriya, U., and Pope, G.A. 2010. Trushenski, S.P. 1977. Micellar flooding: sulfonate-polymer interaction. In
The Role of Co-Solvents and Co-Surfactants in Making Chemical Improved Oil Recovery by Surfactant and Polymer Flooding, ed. D.O.
Floods Robust. Paper SPE 130007 presented at the SPE Improved Oil Shah and R.S. Schechter, 555–575. New York: Academic Press.
Recovery Symposium, Tulsa, 24–28 April. doi: 10.2118/130007-MS. Uren, L.C. and Fahmy, E.H. 1927. Factors Influencing the Recovery of
Salager, J.L., Bourrel, M., Schechter, R.S., and Wade, W.H. 1979. Mix- Petroleum from Unconsolidated Sands by Waterflooding. SPE-927318-
ing Rules for Optimum Phase-Behavior Formulations of Surfactant/ G. Petroleum Transactions, AIME, 77: 318–335.
Oil/Water Systems. SPE J. 19 (5): 271–278. SPE-7584-PA. doi: Wagner, O.R. and Leach, R.O. 1959. Improving Oil Displacement Effi-
10.2118/7584-PA. ciency by Wettability Adjustment. SPE-1101-G. Trans., AIME, 216:
Salter, S.J. 1977. The Influence of Type and Amount of Alcohol on 65–72.
Surfactant-Oil-Brine Phase Behavior and Properties. Paper SPE Wang, D., Cheng, J., Yang, Z., Li, Q., Wu, W., and Yu, H. 2001.
6843 presented at SPE Annual Meeting, Denver, 9–12 October. doi: Successful Field Test of the First Ultra-Low Interfacial Tension
10.2118/6843-MS. Foam Flood. Paper SPE 72147 presented at the SPE Asia Pacific
Schechter, D.S., Zhou, D., and Orr, F.M. Jr. 1994. Low IFT drainage Improved Oil Recovery Conference, Kuala Lumpur, 6–9 October.
and imbibition. J. Pet. Sci. Eng. 11 (4): 283–300. doi: 10.1016/0920- doi: 10.2118/72147-MS.
4105(94)90047-7. Wang, F.H.L. 1993. Effects of Reservoir Anaerobic, Reducing Conditions
Scriven, L.E. 1976. Equilibrium bicontinuous structure. Nature 263 (09 on Surfactant Retention in Chemical Flooding. SPE Res Eng 8 (2):
September 1976): 123–125. doi: 10.1038/263123a0. 108–116. SPE-22648-PA. doi: 10.2118/22648-PA.

906 December 2011 SPE Journal


Wellington, S.L. and Richardson, E.A. 1997. Low Surfactant Concentration George J. Hirasaki had a 26-year career with Shell Development
Enhanced Waterflooding. SPE J. 2 (4): 389–405. SPE-30748-PA. doi: and Shell Oil Companies before joining the chemical engineer-
10.2118/30748-PA. ing faculty at Rice University in 1993. At Shell, his research areas
Wessen, L.L. and Harwell, J.H. 2000. Surfactant Adsorption in Porous were reservoir simulation, enhanced oil recovery, and forma-
Media. In Surfactants: Fundamentals and Applications in the Petroleum tion evaluation. At Rice, his research interests are in NMR well
logging, reservoir wettability, surfactant enhanced oil recovery,
Industry, ed. L.L. Schramm. New York: Cambridge University Press.
foam-mobility control, gas-hydrate recovery, asphaltene depo-
Yan, W., Miller, C.A., and Hirasaki, G.J. 2006. Foam sweep in frac- sition, and emulsion separation. He holds a BS degree from
tures for enhanced oil recovery. Colloids and Surfaces A: Physi- Lamar University and a PhD degree from Rice University, both in
cochem. Eng. Aspects 282–283: 348–359. doi: 10.1016/j.colsurfa. chemical engineering. He received the SPE Lester Uren Award in
2006.02.067. 1989 and was named an Improved Oil Recovery Pioneer at the
Yang, H.D. and Wadleigh, E.E. 2000. Dilute Surfactant IOR—Design 1998 SPE/DOE IOR Symposium. Clarence A. Miller is Louis Calder
Improvement for Massive, Fractured Carbonate Applications. Paper Professor Emeritus of Chemical and Biomolecular Engineering
SPE 59009 presented at the SPE International Petroleum Conference at Rice University and a former chairman of the department.
Before coming to Rice, he taught at Carnegie-Mellon University.
and Exhibition in Mexico, Villahermosa, Mexico, 1–3 February. doi:
He has been a Visiting Scholar at Cambridge University, the
10.2118/59009-MS. University of Bayreuth (Germany), and the Delft University of
Zhang, D.L., Liu, S., Puerto, M., Miller, C.A., and Hirasaki, G.J. 2006. Technology (Netherlands). His research interests center on emul-
Wettability alteration and spontaneous imbibition in oil-wet carbonate sions, microemulsions, and foams and their applications in deter-
formations. Journal of Petroleum Science and Engineering 52 (1–4): gency, enhanced oil recovery, and aquifer remediation. He is
213–226. doi: 10.1016/j.petrol.2006.03.009. coauthor of the book Interfacial Phenomena, now in its second
Zhang, J., Nguyen, Q.P., Flaaten, A.K., and Pope, G.A. 2008. Mechanisms edition. He holds BS and PhD degrees in chemical engineering
of Enhanced Natural Imbibition with Novel Chemicals. Paper SPE from Rice and the University of Minnesota. Maura C. Puerto is
113453 to be presented at the SPE/DOE Symposium on Improved Oil a research scientist at Rice. She retired from ExxonMobil as a
Research Associate after more than 25 years of service in the
Recovery, Tulsa, 20–23 April. doi: 10.2118/113453-MS.
Reservoir Division. She is known as an expert in application of
Zhang, R. and Somasundaran, P. 2006. Advances in adsorption of sur- surfactants to processes involving flow through porous media,
factants and their mixtures at solid/solution interfaces. Adv. Colloid and is skilled in the area of designing laboratory coreflood test-
Interface Sci. 123–126 (Special Issue, 16 November 2006): 213–229. ing and history matching the results by simulation. Before joining
doi: 10.1016/j.cis.2006.07.004. Exxon in 1974, she was employed in quality control and prod-
Zhang, Y., Yue, X., Dong, J., and Liu, Y. 2000. New and Effective Foam uct development for Almay Hypoallergenic Cosmetics and was
Flooding To Recover Oil in Heterogeneous Reservoir. Paper SPE 59367 technical director of the cosmetic branch at the Cuban Institute
presented at the SPE/DOE Improved Oil Recovery Symposium, Tulsa, of Soap and Cosmetics. Since her retirement from Exxon, she has
3–5 April. doi: 10.2118/59367-MS. participated on a part-time basis in research on enhanced oil
recovery and aquifer remediation at Rice. She was the recipi-
Zhao, P., Jackson, A.C., Britton, C., Kim, D.H., Britton, L.N., Levitt, D.B.,
ent in 1980 of the SPE Cedric K. Ferguson Award, and in 1996
and Pope, G.A. 2008. Development of High Performance Surfactants she received, based on her work on CO2 solubility in LVP solvents
for Difficult Oils. Paper SPE 113432 presented at the SPE/DOE for aerosols, the Quality Award from Exxon Chemicals for turning
Symposium on Improved Oil Recovery, Tulsa, 20–23 April. doi: environmental regulations into opportunities. She holds a chemi-
10.2118/113432-MS. cal engineering degree from Oriente University in Cuba.

December 2011 SPE Journal 907


Carbon Dioxide Storage Capacity
of Organic-Rich Shales
S.M. Kang, E. Fathi, R.J. Ambrose, I.Y. Akkutlu, and R.F. Sigal, The University of Oklahoma

Summary geological formation. At a glance, a formation with a high pore vol-


This paper presents an experimental study on the ability of ume would appear to be a good candidate for the purpose. However,
organic-rich-shale core samples to store carbon dioxide (CO2). An this initial impression needs to be tempered because not all high-
apparatus has been built for precise measurements of gas pressure porosity formations are suitable for permanent storage of the gas.
and volumes at constant temperature. A new analytical methodol- Some of them lack a suitable storage environment that will foster
ogy is developed allowing interpretation of the pressure/volume physical mechanisms of gas trapping. In the absence of a trapping
data in terms of measurements of total porosity and Langmuir mechanism, a free-gas cap is artificially created in the formation,
parameters of core plugs. The method considers pore-volume which may not warrant long-term storage of the injected gas. The
compressibility and sorption effects and allows small gas-leakage trapping is associated with fluid/fluid or fluid/solid interactions in
adjustments at high pressures. Total gas-storage capacity for pure porous media, such as dissolution, physical adsorption, or some
CO2 is measured at supercritical conditions as a function of pore homogeneous and heterogeneous reactions. Depleted oil reservoirs,
pressure under constant reservoir-confining pressure. It is shown for example, could be considered among the acceptable target loca-
that, although widely known as an impermeable sedimentary rock tions. The trapping mechanism is mainly driven by absorption of the
with low porosity, organic shale has the ability to store significant injected gas by the immobile residual oil. This is often accomplished
amount of gas permanently because of trapping of the gas in an through multiple-contact (dynamic) miscibility, which involves
adsorbed state within its finely dispersed organic matter (i.e., kero- simultaneous phase-change and mass-transport phenomena. These
gen). The latter is a nanoporous material with mainly micropores reservoirs can also trap CO2 that is not dissolved in the reservoir
(< 2 nm) and mesopores (2–50 nm). Storage in organic-rich shale fluids to the extent that the reservoir seals for hydrocarbons also
has added advantages because the organic matter acts as a molecu- form a capillary seal for CO2. Saline aquifers, on the other hand,
lar sieve, allowing CO2—with linear molecular geometry—to allow aqueous-phase precipitation reactions as well as absorption by
reside in small pores that the other naturally occurring gases cannot the formation water. However, the dissolved gas promotes density-
access. In addition, the molecular-interaction energy between the driven natural convection of water and the related hydrodynamic
organics and CO2 molecules is different, which leads to enhanced instabilities; consequently, the injected gas could be transported
adsorption of CO2. Hence, affinity of shale to CO2 is partly because and dispersed over large distances, leading to uncertainties in its
of steric and thermodynamic effects similar to those of coals that fate. Coalbeds and naturally occurring gas-hydrate reservoirs have
are being considered for enhanced coalbed-methane recovery. been proposed as economically feasible choices of location because
Mass-transport paths and the mechanisms of gas uptake the injected CO2 not only could be sequestered in sorbed states
are unlike those of coals, however. Once at the fracture/matrix (adsorbed on microporous coal material surfaces and absorbed into
interface, the injected gas faces a geomechanically strong porous organic macromolecular openings in coal and in water) but also
medium with a dual (organic/inorganic) pore system and, there- could participate in enhanced recovery of natural gas by an in-situ
fore, has choices of path for its flow and transport into the matrix: molecular swapping mechanism that promotes release and displace-
the gas molecules (1) dissolve into the organic material and diffuse ment of methane as a free gas. Introduction of greenhouse gases into
through a nanopore network and (2) enter the inorganic material these formations, however, is a difficult field operation because of
and flow through a network of irregularly shaped voids. Although a significant loss in well injectivity or no injectivity.
gas could reach the organic pores deep in the shale formation fol- The present work is a fundamental-level investigation on
lowing both paths, the application of the continua approximation gas shales as another location for subsurface CO2 sequestration.
requires that the gas-flow system be near or beyond the percola- Thermally mature organic-rich shales have proved to be prolific
tion threshold for a consistent theoretical framework. Here, using reservoirs for natural-gas production. In the United States, there is
gas permeation experiments and history matching pressure-pulse large production from numerous wells and an extensive pipeline
decay, we show that a large portion of the injected gas reaches the structure associated with organic-shale-gas reservoirs located near
organic pores through the inorganic matrix. This is consistent with major population centers, such as the Barnett, adjacent to Dallas/
scanning-electron-microscope (SEM) images that do not show Fort Worth, and the Marcellus, in New York state. These shales
connectivity of the organic material on scales larger than tens of store methane over geologic time spans both in sorbed states and
microns. It indicates an in-series coupling of the dual continua in as free gas. Also, much like the other shales playing the roles of a
shale. The inorganic matrix permeability, therefore, is predicted to barrier or seal in a petroleum-reservoir system, they have ultralow
be less, typically on the order of 10 nd. More importantly, although permeability. Therefore, as the gas wells are depleted, these shales
transport in the inorganic matrix is viscous (Darcy) flow, transport will become a natural candidate to sequester CO2.
in the organic pores is not due to flow but mainly to molecular Shale sediments with potential for natural-gas production are
transport mechanisms: pore and surface diffusion. generally rich in organic matter, also known as kerogen. Table 1
shows that the total organic content (TOC) in shale could be up
Introduction to 10% of the total shale weight for the North American shale-gas
One of the primary considerations in subsurface sequestration of plays. Although this may not be perceived as a large value, the
anthropogenic CO2 is the knowledge of the gas storability of the significance of the organic matter in the storage of gases in shale
becomes more obvious when one considers its contribution to the
total shale pore volume. In the unconventional-gas-research com-
munity, it is increasingly accepted that the organic matter is the
Copyright © 2011 Society of Petroleum Engineers main constituent of total pore volume, which is associated with
This paper (SPE 134583) was accepted for presentation at the SPE Annual Technical in-situ generation and storage of natural gas. In this study, we
Conference and Exhibition, Florence, Italy, 20–22 September 2010, and revised for experimentally predict an organic-pore-volume/total-pore-volume
publication. Original manuscript received for review 28 June 2010. Revised manuscript
received for review 28 September 2010. Paper peer approved 5 October 2010. ratio up to 70% for Barnett-shale samples. The organic matter in
shale could be a suitable place for CO2 sequestration.

842 December 2011 SPE Journal


into the pulse-decay analysis through forward simulation of gas
TABLE 1—TYPICAL TOC
OF NORTH AMERICAN SHALE-GAS PLAYS* flow and transport in the core plug. We predict shale-gas quanti-
ties of practical interest, such as organic pore volume, porosity,
Shale or Play Average TOC (wt%) permeability, and molecular-diffusion coefficients, as a function
of pore pressure. These measurements will allow us to investigate
Barnett 4
the nature of gas transport in organic-rich shale.
M arcellus 1–10
Haynesville 0-8 Pore-Scale Considerations in Organic-Rich
Horn River 3 Gas Shale
W o o d f o rd 5 Fig. 1a shows a typical 2D focused-ion-beam/SEM (FIB/SEM)
* Ambrose et al. 2010
image of gas shale at micron scale. The image shows organic
matter in dark gray as a finely dispersed porous material imbed-
ded within light-gray inorganic clays. Pores are shown in black,
and note that most of them are within organic islands, or kerogen
The paper is organized as follows. At micron scale, the pockets, that have characteristic size between 200 and 500 nm.
organic-rich gas shales have been visually investigated recently The average size of the organic pores is typically much smaller
by several groups (Loucks et al. 2009; Wang and Reed 2009; than those we observe elsewhere in the inorganic matrix. In the
Sondergeld et al. 2010). Salient results relevant to sequestration next section, we present an analytical approach showing that the
will be summarized from these petrophysical investigations. Fol- organic material mainly consists of micropores (pore lengths less
lowing that, we introduce an approach that uses typical equilibrium than 2.0 nm) and mesopores (pore lengths between 2 and 50 nm),
adsorption data, assumes Langmuir monolayer adsorption, and with an average pore size below 4–5 nm. Unfortunately, this is a
yields effective organic pore size as an indication of organic pore length-scale that is too close to the molecular realm and beyond the
space available for the storage. It is shown that the shale organics maximum resolution of the current SEM technology. This suggests
are nanoporous materials. Although they may not be suitable for that, if one were allowed to look carefully at the fine details of
storing large volumes of CO2 as free gas, they are characterized the pore structure in the organics in Fig. 1, uniformly distributed
by a large internal surface area suitable for trapping significant micropores would have appeared, surrounding the currently visible
amounts of gas at an adsorbed state. Having small organic ones, giving a sponge-like appearance to the kerogen pockets.
pores introduces new complications into the storage-capacity Ambrose et al. (2010) recently reported the results of an
and transport measurements, however. The organic pore volume analysis involving hundreds of 2D images of organic shale samples,
is a dynamic stress-dependent quantity. Next, in the light of this sequentially obtained using ion milling, and their recombined 3D
observation, we introduce a new methodology for the measure- shale segmentations. The digital segment analysis reveals kerogen
ment of total shale porosity and gas-storage capacity where the pore networks. A typical network consisting of large interconnected
pore volume is allowed to change as a function of pore pressure pockets of kerogen is shown in Fig. 1b. Although the samples repre-
because of both adsorption/absorption and because of pore-volume sent an extremely small portion of the reservoir and additional stud-
compressibility. The method is based on a five-parameter nonlin- ies are currently necessary for statistical inference, the following
ear model of storability under equilibrium conditions, where the fundamental conclusions will be made for the consideration of CO2
parameters will require a minimum five-stage laboratory-measure- storage and transport: The organic-rich-shale matrices consist of
ment process using the hydrocarbon gas in the reservoir or CO2. organic and inorganic materials that could be dispersed within each
The method will be applied to the estimation of storage with pure other or bicontinuous and, at times, intertwined; a major fraction of
methane and pure CO2. Finally, shale-gas-permeation experiments the total porosity within the system is associated with the kerogen
and pressure-pulse-decay measurements are performed, investi- network, and therefore much of the gas-storage capacity within
gating mass-transport mechanisms in the presence of adsorption shale is predominantly associated with the organic fraction of the
using various gases. Although pulse-decay analysis is common in rock matrix. In the following pages, we will take advantage of these
the industry for the laboratory measurement of low-permeability observations and develop a new conceptual shale-matrix model that
samples, our approach is new and unconventional. It involves a involves a porous medium with dual (organic/inorganic) -porosity
simulation-based nonlinear history-matching algorithm where the continua. The approach will allow us to quantify and measure total
new pore-scale considerations of gas shales have been incorporated gas storage and transport in organic-rich shale samples.

(a) (b)

Fig. 1—(a) 2D backscatter FIB/SEM image of a gas-shale sample. (b) 3D FIB/SEM shale segmentation. The shale segmentation
size is 4 µm high, 5 µm wide, and 4 µm deep.

December 2011 SPE Journal 843


Effective Organic Pore Size. An effective-pore-size-estimation to cover for a uniform size sphere and cylinder. In the case of a
approach considers that gas adsorption takes place only on the cylindrical pore, we will assume that it is a cylinder without any
organic pore walls in the kerogen pockets. The approach is based flat ends so only the curved surface area will be calculated. For
on the premise that an estimate of total surface area of the organic a sphere, the volume is V = 4r3/3 and surface area is As = 4r2.
material in shale can be realized if a method is developed whereby Hence, for the spherical pore, the volume/surface-area ratio would
the amount of adsorbed gas corresponding to monolayer coverage be V/As = (4r3/3)/(4r2) = r/3. Using similar arguments, it could
can be determined. Physically, the monolayer approach may not be shown that the ratio for a cylindrical pore is equal to r/2. Next,
mean much when pores are so small, although a simple surface- we calculate the effective pore sizes corresponding to spherical and
science concept could help us illustrate the potential features of cylindrical geometries using the gas-filled organic pore volume of
organic shale. For simplicity, we consider that all of the pores in 5.8967 × 1024 nm3 (or 5,896.7 cm3):
the organics will have the same shape: either spherical or cylindri-
cal because both shapes will be investigated. We used methane for ⎛V⎞ ⎛ 5.8967 × 10 24 nm 3 ⎞
rsphere = 3 ⎜ ⎟ = 3 ⎜ ≅ 3.0 nm .
the analysis, and the shale has the physical characteristics shown ⎝ As ⎠ ⎝ 5.7562 × 10 24 nm 2 ⎟⎠
in Table 2. First, the total volume available for gas sorption must
be calculated. It is assumed that the pores containing water will ⎛V⎞ ⎛ 5.8967 × 10 24 nm 3 ⎞
rcylinder = 2 ⎜ ⎟ = 2 ⎜ ≅ 2.0 nm .
not have any surfaces where gas adsorption can occur and that ⎝ As ⎠ ⎝ 5.7562 × 10 24 nm 2 ⎟⎠
the volume of gas that can dissolve in the clay-bound water is
negligible. Hence, total shale gas pore volume per bulk volume Table 3 shows the estimated values of spherical and cylindri-
is equal to  (1−Sw). Consider that only k0 fraction of this total cal pore radii in the organics as a function of k0. Assuming slit-
gas pore volume is associated with the organics; hence, the gas type pore geometry, Schettler et al. (1989) estimated an average
pore volume in the organics is k0  (1−Sw). The expression can be pore diameter of 5.5 nm for Devonian shales using the Brunauer-
written in terms of volumes (cm3) as follows: [mass/density]rock k0 Emmett-Teller multilayer-adsorption isotherm method. This is a
 (1−Sw). Thus, using the values in Table 2, the gas-filled organic same-order-of-magnitude value. On the basis of the calculations,
pore space for 1.0 ton of shale rock is equal to it can be argued that a significant portion of the organic pores are
small (less than 4–6 nm) and, therefore, may not be accessible
⎛ 1 ton ⎞ ⎛ 907,184.75 g ⎞ with the resolution of the SEM image shown in Fig. 1. These small
⎜ 2.5g cm 3 ⎟ × ⎜⎝ ⎟⎠ organic pores, with their large surface area, are an ideal place for
⎝ ⎠ ton
trapping gas in an adsorbed state and for its long-term storage.
×0.55 × 0.05 × (1 − 0.35) = 5, 896.7 cm 3. However, they make the measurements and analyses of gas storage
and transport somewhat complicated because, at that length scale
The number of lbm mol in 1.0 scf of gas is • The thickness of the adsorption layer becomes comparable
n P to the organic pore size. Under thermodynamic equilibrium (no
= std flux of mass or energy), some of the gas molecules inside the
V RTstd
organic pore are closer and, therefore, under stronger influence
14.69 psi of the pore walls. These molecules have limited mobility and
=
⎛ scf psi ⎞ relatively low kinetic energy. They participate in the formation of
⎜⎝ 10.7316 ⎟ × ( 60 + 459.67 ) °R
°R lbm mol ⎠ a dense (liquid-like) adsorption layer that covers internal surfaces
lbm mol of the pore walls; see Fig. 2a. Next to the adsorption layer, we
= 2.6340 × 10 −3 . typically observe phase-transition layers where the molecules are
scf constantly being adsorbed and desorbed under equilibrium. The
Therefore, the number of methane molecules in 50 scf/ton of gas molecules in phase transition are relatively less dense and more
(Langmuir volume) would be mobile with some kinetic energy, although they are under some-
what reduced influence of the walls. The rest of the gas molecules
(50.0 scf ) × ⎛⎜⎝ 2.6340 × 10 −3
lbm mol ⎞ ⎛ g ⎞ are at the central portion of the pore (if any space is left), and they
⎟⎠ × ⎜⎝ 453.5924 ⎟
scf lbm ⎠ are not under the influence of pore walls. They are the free-gas
molecules, with a density equal to the fluid bulk density in the
⎛ molecules ⎞ absence of pore walls. They mainly interact among themselves and
× ⎜ 6.0221 × 10 23 = 3.5976 × 10 25 molecules.
⎝ g mol ⎟⎠ with the gas molecules in phase transition. In the case of methane
adsorption in a 3.6-nm-thick organic slit pore shown in Fig. 2, in
If one adsorbed methane molecule covers 0.16 nm2 (Anderson and total 2 × 0.37 = 0.74 nm of space next to the walls will be occupied
Pratt 1985), assuming monolayer coverage, we can estimate the by the adsorption layer, thus leaving only 3.6−0.74 = 2.86-nm-
total area covered by this number of methane molecules as thick pore space available for the phase-transition molecules and
0.16 nm 2
3.5976 × 10 25 molecules × = 5.7562 × 10 24 nm 2.
molecule
TABLE 3—ESTIMATED ORGANIC PORE RADII
This is a significant value, indicating that the organic pores have FOR VARYING k0 VALUES
relatively large internal surface area. We can use the ratio of vol-
ume to surface area to calculate the radius the molecules needed Fraction of Porosity Spherical Pore Cylindrical Pore
in Organic ( k0) Radius, rsphere (nm) Radius, rcylinder (nm)

0.1 0.61 0.41


0.2 1.23 0.82
TABLE 2—SHALE PARAMETERS
FOR EFFECTIVE-PORE-SIZE CALCULATIONS 0.3 1.84 1.23
0.4 2.46 1.64
Parameter Symbol Value
0.5 3.07 2.05
Total porosity φ 5% 0.6 3.69 2.46
Organic/total pore volume ratio k 0.5 0.7 4.30 2.87
Water saturation Sw 35% 0.8 4.92 3.28
3
Bulk density 2.5 g/cm 0.9 5.53 3.69
Langmuir volume (methane) VL 50 scf/ton 1 6.15 4.10

844 December 2011 SPE Journal


Nonequilibrium

Slippage

0.35 Adsorption layer

0.30
Phase
0.25 Hopping
transition layers
ρCH4, g/cm3

0.20
Free gas
0.15

0.10

0.05

0.00
0.0 3.8 7.6 11.4 15.2 Direction of external molecular flux
Pore Half Length, Å
(a) (b)

J K,x
(c) J S,x
Fig. 2—(a) Molecular layer density for methane at 176°F (80°C) across the half-length of a 3.74-nm organic slit pore. Results
obtained using equilibrium molecular dynamics simulation carried out in the canonical ensemble. The estimated pore pressure
is approximately 2,000 psi. (b) Schematic representation of gas slippage and hopping mechanisms of molecular transport. (c)
Schematic of gas mass transport in organic nanopores. The arrows represent diffusive mass fluxes of the free-gas phase (thick
arrow) and the adsorbed phase (thin arrow) in the x direction. JK,x and JS,x are the contributions of free-gas pore diffusion and
adsorbed-phase surface diffusion, respectively.

for the free-gas molecules. Density structure of gas molecules in the gas mass transport in the organics should be in the transitional
small organic pores and its effect on shale-gas in-place estima- slippage-flow regime. This means that the assumption of Darcy
tions have been studied recently by our group; see Ambrose et al. flow is invalid in the organic pores in shale. Therefore, a new
(2010). approach is necessary for modeling the gas mass transport, which
• The mean free path of gas molecules (i.e., average distance will incorporate non-Darcy transport of the free-gas molecules.
gas molecules travel between two successive collisions) becomes • Adsorbed-phase transport could be another important mass-
comparable to the available pore space. Consequently, the free-gas transfer mechanism. Fig. 2c shows the free-gas and adsorbed-phase
molecules interact vigorously with the molecules in phase transi- components of total mass flux in the organic micro- and mesopore.
tion and with those that make up the adsorption layer and pore wall, Under the influence of an external flux, some of the molecules in
compared with the level of interaction the free-gas molecules have the adsorbed-phase and those in phase transition can overcome the
among themselves at the central portion of the pore. When thermo- local interactions with the wall and develop a hopping mechanism.
dynamic equilibrium is perturbed because of an externally applied (Much like drifting of a tumbleweed or tree leaves on the ground
molecular flux, for example, these amplified interactions may lead under the influence of a mild wind.) This transport mechanism, also
to kinetic effect that causes streaming of the gas molecules—in known as surface diffusion, is not important if it takes place in a
particular, those in phase transition. This is the occurrence of the large pore with an almost infinite number of free-gas molecules;
so-called slippage effect (or molecular streaming) in the organic but, if the discussion is on nanopores with a large surface area, it
pore. It can cause the collapse of laminar-flow (parabolic) -veloc- may easily contribute to the total mass flux and, hence, enhance
ity profile with zero velocity at the wall. Consequently, the free- the kinetic effect by the pore walls. Let us reconsider the layers of
gas mass transport would show deviations from Darcy’s law and molecules indicating nonuniform structure (damped oscillation) in
would be represented better with molecular (pore) diffusion. A methane density in a 3.6-nm organic pore in Fig. 2. It is not dif-
useful dimensionless parameter in the literature to determine the ficult to imagine that the molecular layers corresponding to free gas
changes in flow regime of a gas system is the Knudsen number would not develop and all of the existing molecules would be either
(NKn). This number is the ratio of the molecular mean free path to adsorbed or under the influence of pore walls if the pore size were
the pore length scale. When NKn > 0.1, the kinetic effect becomes less then 2×1.08 = 2.16 nm. Then, the only transport mechanism
dominant by the wall in the phase-transition region. Further dis- during the gas release from or uptake by the pore would be because
cussion on the flow regimes can be found in Roy et al. (2003). of surface diffusion. We note that, although they both take place in
In Fig. 3, we consider methane existence in organic pore volume the phase-transition region, surface diffusion is a different transport
with characteristic pore length varying between 2 and 10 nm and mechanism than the slippage; at a given pressure and temperature,
estimate the NKn for varying pore pressures. The values inside the surface diffusion is driven by gradients in adsorbed-phase density,
ellipse (typical initial shale-gas-reservoir conditions) indicate that whereas slippage is driven by the pore size.

December 2011 SPE Journal 845


100.00
2-nm pore radius
10-nm pore radius
10.00

N Kn
1.00

0.10

0.01
10 100 1,000 10,000
Pressure, psia

Fig. 3—Knudsen number vs. pore pressure for pore-throat radii calculated from the Langmuir isotherm at reservoir temperature
of 188°F.

• The assumption of continuum-fluid mechanics may −2r 8 RT ⎛ dn ⎞


become questionable—in particular, in the micropores. A 2.86- J K ,x = ⎜ ⎟ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
3  M ⎝ dx ⎠
nm pore can, at most, include only seven methane molecules in a
row across the pore space. Similarly, approximately 30 methane Equivalently, this expression can be reformulated in terms of
molecules would fit in a 10-nm organic pore. Hence, the number moles of concentration C rather than molecular concentration:
of molecules is limited inside the organic pores, which may not
allow the assumption of a representative elementary volume −2r 8 RT ⎛ dC ⎞
J K ,x = ⎜ ⎟ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
for the fluid (i.e., the classical fluid-particle concept of fluid 3  M ⎝ dx ⎠
mechanics) with average intensive properties, such as density
and viscosity. By analogy to Fickian gas diffusion, we now define a molecu-
In this work, these pore-scale considerations will be part of lar-diffusion coefficient for transport in the capillary as
the gas-storage and -transport measurements. Gas-storage mea-
surements do not necessarily require the assumption of continuum, 2r 8 RT
DK ,capillary = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
although the transport calculations do. Our group has other investi- 3 M
gations based on noncontinuum approaches (e.g., molecular mod-
eling and molecular/continuum multiscale modeling), although Thus, the diffusion coefficient is proportional to the effective
they are discussed elsewhere; see Diaz-Campos et al. (2009) and (adsorption layer-corrected) pore radius and the mean molecular
Ambrose et al. (2010). During the measurements at varying pore velocity. One can relatively easily obtain typical values to compare
pressures, the organic pore volume available for the free gas will with the experimentally obtained organic pore diffusion coeffi-
be adjusted for the isothermal pore compressibility effect (effective cient. For the purpose, we consider a 1.5-nm cylindrical pore and
stress) and also for the presence of an adsorption layer. During perform the calculations at our laboratory conditions:
transport calculations, we will differentiate free-gas and adsorbed-
phase diffusive transport in the organics from gas flow in inorganic DK ,capillary =
−9
2r 8 RT 2 × 1.5 × 10 m
=
( )
matrix and in fractures using a separate transport mechanism for 3 M 3
each. In the organics, diffusion of the adsorbed phase accounts for
⎛ kg m 2 ⎞
surface diffusion effects of molecules in phase transition, whereas 8 × ⎜ 8.3145 2 × ( 298.15 K )
diffusion of free molecules represents transport in the pore space ⎝ s K mol ⎟⎠
×
and slippage in the phase-transition region. ⎛ kg ⎞
3.1416 × ⎜ 16.042 × 10 −3 ⎟
⎝ mol ⎠
Interpretation of Organic Pore Diffusion. Consider flux of
free-gas molecules across the available organic pore space in the m2 cm 2
= 6.3 × 10 −7 = 6.3 × 10 −3 .
direction parallel to its inlet and outlet ends. The net free-gas mass s s
flux is proportional to the difference in gas number densities n1
One would expect this estimate to be somewhat larger than the
and n2 at the two ends: JK = wc(n2− n1), where w is a dimension-
experimentally obtained molecular diffusivity because the mea-
less probability factor and c is the mean molecular speed. The
surement is the outcome of an expression derived for a single
mean molecular speed could be estimated from the kinetic theory
capillary pore. Typically, the two can be roughly related through
of gases using c = [8 RT ( M )] , where M is the gram molecular the porosity/tortuosity ratio as
weight, T is the absolute temperature, and R is the universal gas
constant (Roy et al. 2003). The probability factor is dependent on ⎛ ε ⎞⎛  ⎞
DK = ⎜ k ⎟ ⎜ 2 ⎟ DK ,capillary, . . . . . . . . . . . . . . . . . . . . . . . . . (4)
the geometry of the pore and requires knowledge of the appropri- ⎝ ε TOC ⎠ ⎝  ⎠
ate scattering law. Considering diffusive scattering and simple
geometry, such as a capillary with a length L and radius r, the where TOC is the kerogen/total-volume ratio and  is tortuosity.
value of w for the condition L >> r is equal to 2/3(r/L). Substitut- Because it cannot be measured directly, there exists uncertainty in
ing these expressions into the flux equation so that it gives mass the values of tortuosity of nanoporous materials. Values between
flux along the axis of a relatively long capillary pore and writing 1.5 and 8 have been reported by Wei et al. (2007) and Busch et al.
in differential form using the convention that the molecules move (2008). If the kerogen porosity (the first parenthesis term on the
from high to low concentration of molecules, we obtain right-hand side of Eq. 4 multiplied with ) is 20%, and a kerogen

846 December 2011 SPE Journal


tortuosity of 3.0–6.0 is assumed, one would expect that the organic sure vessel is plumbed so that the sample can be kept at a confining
pore-diffusion coefficient measured using pressure-pulse decay DK pressure that corresponds to the confining stress that the sample is
should be on the order of 10−5 cm2/s (i.e., two orders of magnitude exposed to in the reservoir. During the measurements, the confin-
smaller). In addition, the adsorbed-phase transport can be captured ing pressure is always kept greater than the pore pressure.
similarly using Fick’s law: The storability measurement is a multistage measurement.
Each stage starts with Valve 2 closed. To start, Volume 1 is raised
⎛ dC ⎞ to a pressure P1, and then Valve 1 is closed. The sample and dead
J S ,x = − DS ⎜  ⎟ , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
⎝ dx ⎠ Volume 2 are at an initial starting pressure P2. The test consists of
opening Valve 2 and recording the pressure as a function of time
where Cµ is the adsorbed-phase amount in terms of moles per solid until it has stabilized at a new pressure Pf . The pressure-decay
organic volume and DS is the porosity- and tortuosity-corrected curve can be used to obtain the transport coefficients. The test can
surface diffusivity. Typically, for Cµ, we would expect one-order- be run with P1 less than P2 (for the gas-release measurements) or P1
of-magnitude-smaller values then the free-gas amount C, although greater than P2 (for the gas-uptake measurements). In the first case,
its gradient could vary and be relatively large during the production the decrease in storability with decrease in pressure is measured;
and sequestration operations. and, in the second case, the increase in storability is measured.
When evaluating the capacity of a shale reservoir to store For the second case, the analysis of a given stage proceeds
CO2, an essential first step is to evaluate the total porosity and as follows. The moles of gas n1 in Volume 1, V1, at pressure P1
how much gas per unit weight (scf/ton) the porous material of the is given by
reservoir can store. In general, this storage capacity is a function
1 1 ( z1 RT ), . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
n1 = PV
of the porous-material properties, the reservoir pore pressure, and
the temperature. When multiple storage mechanisms are present, it
is difficult to characterize the total porosity and storage. First, we where R is the universal gas constant, T is the absolute temperature,
explain a multistage gas-uptake process that allows determination and z1 is the correction factor that accounts for deviation from the
of the total storage capacity and total porosity as a function of ideal-gas law. The correction factor depends on temperature, pres-
pressure and temperature and the parsing of the storage into the sure, and the type of gas. Numerous tables and programs exist that
various storage mechanisms. This measurement can also be taken provide accurate values of z1. Similarly, the moles of gas contained
under in-situ subsurface conditions. in V2 initially are given by
Measurement of Shale-Gas-Storage n2 = P2V2 ( z 2 RT ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
Capacity
For a shale sample, total porosity is defined as the volume of flu- After Valve 2 is opened and the pressure has stabilized at Pf ,
ids (i.e., liquid and free gas, not including adsorbed or absorbed the moles in V1 and V2 are given by
gas) contained in the sample. Hence, the storage capacity of shale
should be divided up into the sorbed component and the pore-vol- n1 f = Pf V1 ( z RT )
f
ume component. In this work, an isothermal multistep gas-uptake
process measures the storage capacity. The minimum number of and
steps required is discussed later.
Fig. 4 shows the basic components needed for the mea- n2 f = Pf V2 ( z RT ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
f
surement. For temperature control, or to make measurements at
reservoir conditions, the apparatus can be enclosed in an oven to Mass balance requires that the decrease of moles of gas in V1,
provide a constant high-temperature environment. Currently, we n1, is equal to the increase in moles of gas in V2, n2, plus the
have two different storability-measurement implementations in the increase in moles of gas stored in the organic pores of the shale
laboratory, one in an environmental chamber and one that operates sample ns. From the previous equations,
at ambient temperature in the laboratory. The apparatus consists
of a pressure vessel that contains the core sample to be tested V1 ⎛ P1 Pf ⎞
n 1 = − ,
along with a dead volume, Volume 2; a gas reservoir, Volume 1; RT ⎜⎝ z1 z f ⎟⎠
a valve, Valve 2, to separate the reservoir volume from the sample
container; a gas source, typically a pressure tank; as valves, Valve V2 ⎛ Pf P2 ⎞
n2 = − ,
1, to isolate the rest of the system from the gas tank; and pressure RT ⎜⎝ z f z2 ⎟⎠
gauges. Not shown are pumps to raise the gas from the bottle
pressure to the needed reservoir pore pressure and the pump and and
plumbing to provide confining pressure on the sample. The pres-
ns = n1 − n2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)

The additional moles of gas ns stored in the sample provide the
Valve 1 increase in storability as a function of the change in pressure.
Reservoir When this process is iterated over multiple steps, it provides the
Volume 1 storability as a function of pore pressure. This is the basic informa-
tion needed to characterize the reservoir and is an essential input
into simulations that model the way gas will be produced from the
reservoir or be sequestered in it. It should be noted that, because
there can be hysteresis in the storability curves, both the increasing
Valve 2 and decreasing storability cases may need to be measured.
Gas To separate the storage in the rock system into the adsorbed/
Source absorbed component and the component stored as free gas in the
pore space, a model of the storage mechanisms needs to be used.
Volume 3 = Volume 2 The moles of gas nsp stored in the sample pore space Vp in terms
+ Sample Volume of the sample pore pressure Ps (Ps equals P2 or Pf) satisfies the
gas law. Therefore,

Fig. 4—Diagram of gas-storage-capacity measurement method. nsp = PsVp ( zs RT )

December 2011 SPE Journal 847


and process using the adsorbed gas or as a separate process using
⎛V P V P ⎞ helium, which is not adsorbed (so only a simple two-stage process
nsp = ⎜ pf f − p 2 2 ⎟ RT . . . . . . . . . . . . . . . . . . . . . . . . (10) needs to be performed). In either case, the pore-volume compress-
⎝ zf z2 ⎠ ibility is defined by Cp so that the volume increase because of
In general, Vp is a function of the pore pressure and the gas increasing pore pressure is given by
type. As we will see, it may not be a single-valued function. The
dependence of Vp on pressure needs to be included in the model Vpp = CpVp0 (Pf –P2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (16)
and will be discussed after the discussion of the adsorption/absorp-
tion term. Vp2 and Vpf are the pore volumes at pressures P2 and Pf , The total change in pore volume going from pore pressure
respectively. P2 to Pf is given by
For both the shale-gas and the coalbed-methane cases, the
amount of gas adsorbed/absorbed by the rock matrix is generally Vp = Vpp+Vpa, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (17)
assumed to be described by a Langmuir isotherm. It is known that
this is only an approximation and other more-exact models could so that
be used to model the adsorption/absorption component of the
storage. For the Langmuir model, all the nonfree gas is assumed Vpf = Vp2+Vp. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (18)
to be adsorbed. The Langmuir isotherm is parameterized by two
quantities, the total moles of adsorbed gas at infinite pore pressure Here, the second-order effect of Cp being dependent on pore
Samax and the Langmuir pressure PL. pressure is ignored. This model of storability has five parameters
The Langmuir equation gives the adsorbed gas storage Sa as to determine, Samax, Vamax, PL, Vp0, and Cp. Simultaneous estima-
a function of the pore pressure P as tion of these quantities will require, at a minimum, a five-stage
measurement process if they are all determined by measurements
Sa =
(
Sa max P PL ), . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (11) using the hydrocarbon gas in the reservoir or CO2. Measurements
1 + P PL with helium could be used to determine Cp and an approximate
value for Vp0, assuming the density of the adsorbed-state gas is
so that nsa is given by constant, and then the measurements could be reduced to two-stage
measurements. The second option is experimentally easier but
⎡ Pf PL P2 PL ⎤
nsa = Sa max ⎢ − would not provide any test of the model assumptions. Combining
( P2 PL ) ⎥⎥⎦
. . . . . . . . . . . . . . . . (12)
⎢⎣ 1 + ( Pf PL ) 1 + the second option with a few detailed multistage measurements
to obtain model-independent storability to test against the model
and might prove to be a cost-effective option.
ns = na + nsp. Multiscale Characterization and
Measurement of Shale-Gas-Transport
The pore volume can change as a function of pressure because
of both adsorption/absorption and because of pore-volume com- Mechanisms
pressibility. Adsorption can affect pore volume by the adsorbed Our arguments based on the indirect laboratory observations
molecules occupying pore space that otherwise would be occupied (SEM images and storage-capacity measurements) suggest that a
by the free gas or by the adsorbed gas closing off pore throats so flow model for shale gas should include dispersed organic porous
gas cannot get through to the larger pore. Absorption can cause the material within the inorganic matrix. A conceptual model showing
matrix to swell or, by swelling, to close off access to pores. The the organic- and inorganic-material distribution in shale is given
effects would look very similar for both adsorption and absorption. in Fig. 5. Note that, compared to coal, the main difference in
Using the Langmuir model associates all the pore-volume changes shale-gas production is that, before reaching the fractures, a large
with adsorption. The Langmuir model can be used to calculate the portion of shale gas in place should be transported first inside
pore-volume loss because of adsorbed gas. This calculation does the kerogen network imbedded in the matrix. It should be noted
not account explicitly for pore-blockage effects, but, to the extent that a large contrast in length scales exists between gas flowing
that a Langmuir isotherm successfully fits the gas adsorption, it is in the organic pores, outside of them in the matrix, and in the
implicitly including these effects. In the Langmuir equation, the fractures. This requires a multicontinuum modeling approach
term (P/PL)/(1+P/PL) represents the fraction of the moles adsorbed involving separate but coupled hydraulic components of the porous
at any pressure. If Vamax represents the maximum adsorbed-gas medium, where each component (organic, inorganic, and fracture)
volume and Va the adsorbed-gas volume at any pressure, then,
assuming Langmuir adsorption,

Va =
(
Va max P PL ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (13)
Organic material
Inorganic
1 + P PL

If Vp0 is the pore volume at zero pore pressure, then the


decrease in Vp at pressure P because of adsorption is given by
Vp0−Vamax (P/PL)/(1+P/PL). The decrease in Vp in each stage is

⎡ Pf PL P2 PL ⎤
Vpa = −Va max ⎢ −
( P2 PL ) ⎥⎥⎦
. . . . . . . . . . . . . . (14)
⎢⎣ 1 + ( Pf PL ) 1 +

If the molar density of the adsorbed gas ads max is known,


then
Organic porous material Inorganic porous material
Vamax = Samax/ads max. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15) with uniform fracture network with nonuniform fracture network
The last consideration is accounting for pore-volume com- (a) (b)
pressibility. Over the pressure ranges applicable, this should be a
small effect with values typically on the order of 10−6. It can be Fig. 5—Distributions of the organic and inorganic materials in
determined in two ways, as part of the multistage measurement coal and shale.

848 December 2011 SPE Journal


Transport in parallel Transport in series
Adsorption+diffusion
Kerogen
K
e Inorganic

Fracture

Fracture
Inorganic r matrix
matrix o
g Adsorption Viscous
Viscous
flow e + flow
n diffusion

(a) (b)

Fig. 6—Conceptual multicontinuum models for gas flow and transport in organic-rich shale during sequestration.

is distributed continuously in space and holds the fundamental ments of permeability and diffusivity of shale are difficult to
porous-medium conditions specified by Bear and Bachmat (1991). implement because of very low mass fluxes and the extremely
Multicontinuum models can be identified by the number of long time needed to reach the steady-state condition. Therefore,
components characterizing the hydraulic behavior of the system transient methods such as pressure-pulse-decay experiments have
and the type of coupling and mass-exchange terms between the been used for the measurements. This method is much faster in
components. The interaction between the components, defined by comparison to the conventional methods and can be used to mea-
finite mass-exchange terms, affects the overall transport behavior sure permeability valuse as low as 10−9 md (Brace et al. 1968; Ning
of the system significantly, while a single-continuum approach 1992; Finsterle and Persoff 1997; Jones 1997). In this method,
neglects this important effect. There are three different types of a small perturbation is introduced to a shale/gas system under
coupling between the components: parallel, series, and selective thermodynamic equilibrium by rapidly changing the downstream
coupling. In parallel coupling, all specified continua are coupled (or upstream) pressure, typically by an approximate 10% of the
with one another directly, while, in the series coupling, they are equilibrium pressure. Consequently, small pressure gradients are
connected in the order of hydraulic conductivity (Lee and Tan generated across the sample, leading to transport of gas across the
1987). In organic-rich-shale-gas reservoirs, multicontinuum could core plug. During the gas permeation, decline in the upstream or
be defined, at the least, by two matrix continua (i.e., organic and downstream pressure is measured to extract transport properties of
inorganic, along with a fracture continuum). the shale using an analytical or numerical inversion method.
In the Effective Organic Pore Size subsection, we argued that
the gas mass transport in the organics is transitional slippage with Results
adsorbed-phase surface-diffusion effects (i.e., non-Darcy). Model- Table 4 shows results from the models that fit the measured stor-
ing of such transitional flow regimes in the presence of adsorption age data for helium, CO2, and methane with two organic-rich shale
is not straightforward, however. Here, following the work by Fathi samples, Sample 21 and Sample 23. These samples are selected
and Akkutlu (2009, in press), we investigate gas-transport behavior because TOC and mineralogy were known. Because, in the initial
in the kerogen by experimentally quantifying the pore and surface measurement stage, the CO2 passes through a pressure where it
diffusion coefficients DK and DS , respectively. Pore diffusion is the can form a liquid phase, there is considerably more uncertainty in
mechanism of transport for the free-gas molecules occupying the the model fits to the measured storage data. Note the estimated Vp0
available pore space, whereas surface diffusion is the mechanism values. The values are quite different for Sample 23, giving 5.38,
of transport for the adsorbed-phase near the internal surfaces of the 3.95, and 2.89% porosity for the gases, respectively. The model
kerogen. During the production, gas in the kerogen pore network fits to the experimental data are not unique. Assuming that all
is eventually released to the relatively larger pores, voids, micro- reasonable fits to the data have the properties of this fit, the varia-
fractures, and cracks in the inorganic matrix or directly into the tions indicate that methane molecules with spherical molecular
nonuniform fracture network shown in Fig. 5b. In its new environ- geometry cannot have access to some of the micropores that CO2
ment, the mechanism of its transport is primarily viscous (Darcy) molecules with linear molecular geometry have access to, and the
flow and, because of limited surface area, the adsorbed amount is CO2 molecules cannot have access to some of the micropores that
considered negligible. In the case of CO2 sequestration, the same the small helium molecules can access. In essence, the organic
arguments are valid, although in the opposite transport direction material of Sample 23 acts as a molecular sieve for fluids used
(i.e., from the fractures deep into the inorganic shale matrix and during measurements; see Fig. 7. Although this effect creates
into the kerogen). Fig. 6 shows two conceptual models of shale-gas uncertainties during the storage-capacity measurements for Sample
flow and transport that are considered in this study for methane 23, in general, it points to an important CO2-sequestration quality
production and CO2 sequestration. In Fig. 6a, parallel coupling is for the shale-gas formations. It suggests that a large fraction of
shown where organic and inorganic continua are coupled directly the injected CO2 can be stored in small pores in which the other
with fractures. In Fig. 6b, all three continua are coupled in series naturally occurring gases, such as methane and nitrogen, cannot
(i.e., fracture → inorganic → organic). We used these conceptual be stored. Compared to methane, for example, CO2 can access an
models during the development of a mathematical formulation additional 100 × (0.4771−0.345)/0.345 = 38.3% of the organic pore
(a group of coupled partial-differential equations) describing the volume when injected into Sample 23 under the same conditions.
mechanisms of gas storage and transport. The formulation is used to The molecular sieving becomes a secondary effect in the case of
simulate gas-permeation experiments using organic-rich shale core Sample 21, however. In this case, methane and CO2 have access
plugs. Later, the developed simulators are used in the experimental to the same pore space, whereas helium can penetrate into an
investigation of multiscale characterization of transport phenomena additional 7.7% organic-micropore volume.
in organic-rich shales. Further discussion on the modeling and As the consequence of changes in the estimated values of the
simulation of shale-gas flow can be found in Fathi (2010). pore volume Vp0, the predicted gas-sorption parameters in Table 4
are quite different. Effect on the Langmuir adsorption isotherm is
Measurements of Transport Coefficients—Unsteady Gas-Per- shown in Fig. 8. Note that, because the pore-volume estimation for
meation Experiments. Steady-state methods for the measure- Sample 21 is consistent, the predicted methane and CO2 isotherms

December 2011 SPE Journal 849


TABLE 4—MEASURED GAS-STORAGE CAPACITY FOR THE OR GANIC-RICH SHALE SAMPLES USING VARIOUS GASES

Sample 21 S a m p le 2 3

He CO2 CH4 He CO2 CH4

Vp0 (cc) 0.42 0.39 0.39 0.65 0.48 0.35


Cp (psi )
–1 6 6 6 6 6 6
5.416 10 5.416 10 5.416 10 7.05 10 7.05 10 7.05 10
PL, CH4 (psi) using Vpo,He 3,232 1,000
using Vpo,CH4 1,800 2,502
Vamax, CH4 (cm )
3
using Vpo,He 0.4193 0.4056
using Vpo,CH4 0.1962 0.1423
Samax, CH4 (mol) using Vpo,He 0.0072 0.0021
using Vpo,CH4 0.0049 0.0041
3
ads, CH4 (mol/cm ) using Vpo,He 0.017 0.005
using Vpo,CH4 0.025 0.029
PL,CO2 (psi) using Vpo,He 500 655
using Vpo,CH4 500 2610
using Vpo,CO2 500 1410
Vamax,CO2 (cm )
3
using Vpo,He 0.4118 0.4300
using Vpo,CH4 0.3850 0.3200
using Vpo,CO2 0.3850 0.3400
Samax, CO2 (mol) using Vpo,He 0.0240 0.0280
using Vpo,CH4 0.0237 0.0211
using Vpo,CO2 0.062 0.0220
3
ads, CO2 (mol/cm ) 0.065

appear distinctly in Fig. 8a. The isotherms for the other sample at any given pore pressure. For both samples, the total CO2-gas-
are shown in Fig. 8b. Indeed, methane isotherms show that the storage capacity is 4 times greater than the total methane-gas-stor-
gas stored as an adsorbed phase is significantly larger when the age capacity at 2,500 psia. Such a large contrast in methane and
pore volume obtained using helium is included in the analysis. The CO2 storage is important for sequestration, although it cannot be
contrast is even larger in the case of CO2. These results indicate explained by the molecular-sieving arguments related to the organ-
the importance of fluid type used for the adsorption analysis. In the ics of certain types of samples. It can be explained only by changes
rest of the paper, we show only the storage and transport results in the fluid’s internal energy with the changes in the fluid type.
for a particular fluid using a sample pore volume measured with Indeed, the molecular-interaction energies (both among the gas
that fluid. molecules and between the organic pore wall and gas molecules)
Modeled total gas storage is given in Figs. 9 and 10. The are different, which leads to significantly enhanced adsorption in
adsorbed-gas and free-gas amounts stored are shown in Fig. 9a the case of CO2. Molecular-level investigation further comparing
for Sample 21 and in Fig. 9b for Sample 23. Clearly, most of the adsorption of fluids and their mixtures in kerogen in the small
injected CO2 is stored in an adsorbed state. For example, at 2,500 organic pores under reservoir conditions is currently necessary.
psia and for Sample 21, the ratio of adsorbed to free CO2 is approx- Results indicate that adsorption is the dominant mechanism
imately 13:1. For Sample 23, that ratio drops to 13:5, although the during the CO2 storage in organic-rich shale. This may also be the
total amount stored is nearly doubled. Hence, the measurements case for methane, although the dominance is not as amplified and
reveal that the samples have larger adsorption capacity for CO2 distinct as in the case of CO2; see Fig. 10. At 2,500 psia, the ratio

Pure Binary mixture Ternary mixture

Fig. 7—Schematic of molecular sieving of gases in kerogen.

850 December 2011 SPE Journal


CH4, Vp0=0.39 cm3 by CO4 CH4, Vp0=0.35 cm3 by CO4
350 0.0012 800
CH4, Vp0=0.42 cm3 by He CH4, Vp0=0.65 cm3 by He
CO2, Vp0=0.39 cm3 by CO4
0.0025

Adsorbed Gas Storage, SCF/ton

Adsorbed Gas Storage, mol/cm3


CO2, Vp0=0.35 cm3 by CO4

Adsorbed Gas Storage, SCF/ton


CO2, Vp0=0.39 cm3 by CO2 700

Adsorbed Gas Storage, mol/cm3


300 0.0010 CO2, Vp0=0.48 cm3 by CO2
CO2, Vp0=0.42 cm3 by He CO2, Vp0=0.65 cm3 by He
600 0.0020
250
0.0008
500
200 0.0015
0.0006 400
150
300 0.0010
0.0004
100
200
0.0002 0.0005
50 100

0 0.0000 0 0.0000
0 1,000 2,000 3,000 4,000 5,000 0 1,000 2,000 3,000 4,000 5,000
Pressure, psia
(a) (b) Pressure, psia

Fig. 8—Estimated Langmuir isotherms for Sample 21 (a) and Sample 23 (b). The parameters used are given in Table 4. The mark-
ers show the five experimental pressure data points used during the prediction of the gas-storage parameters.

500 500
Adsorbed gas, CO2 Adsorbed gas, CO2 0.0016
450 0.0016 450 Adsorbed gas, CO4
Adsorbed gas, CO4 0.0014
0.0014 Free gas, CO2
400 Free gas, CO2 400

Gas Storage, mol/cm3


Free gas, CO4
Gas Storage, SCF/ton

Gas Storage, mol/cm3


Gas Storage, SCF/ton
0.0012
350 Free gas, CO4 0.0012 350
300 300 0.0010
0.0010
250 250 0.0008
0.0008
200 200
0.0006 0.0006
150 150
0.0004 0.0004
100 100
50 0.0002 0.0002
50
0 0.0000 0 0.0000
0 1,000 2,000 3,000 4,000 5,000 0 1,000 2,000 3,000 4,000 5,000
Pressure, psia Pressure, psia

Fig. 9—Estimated adsorbed- and free-gas amounts for Sample 21 (a) and Sample 23 (b). Solid lines are with CO2, and the dashed
lines are with methane.
Adsorbed Gas Storage/Total Gas Storage

Adsorbed Gas Storage/Total Gas Storage

1.5 1.5

CO2 CO2
1 1

CH 4
CH 4
0.5 0.5

0 0
25 1,025 2,025 3,025 4,025 5,025 25 1,025 2,025 3,025 4,025 5,025
Pressure, psia Pressure, psia

(a) (b)

Fig. 10—Estimated ratio of adsorbed gas to total gas stored in Sample 21 (a) and in Sample 23 (b).

December 2011 SPE Journal 851


1,630
Time, seconds
1,620 0 10,000 20,000 30,000 40,000 50,000
0
1,610
Upstream Pressure, psi

–0.5
1,600
Confining pressure 5,000 psi –1
1,590
–1.5
Dark gray curve: experimental data
1,580 Light gray curve: mean of realizations –2

ln (ΔPD)
1,570 Black curve: 112 realizations
–2.5
1,560 –3

1,550 –3.5

1,540 –4 μgLcg
kg=|slope|
1,530 –4.5 6.805×10–5f1A(1/V1+1/V2)
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
×104 –5
Time, seconds
(a) (b)

Fig. 11—(a) Example history-matching results for Sample 21 using upstream pressure decline. 112 realizations are used to obtain
the optimum values of the transport coefficients shown in the first row in Table 5. (b) Permeability for Sample 21 is obtained
using the pulse decay shown on the left and using Jones’ straight-line method (Jones 1997).

of adsorbed CO2 to total CO2 gas storage is approximately 0.93 nature of multiscale gas transport in shale are discussed for the
for Sample 21; on the other hand, the ratio for methane is approxi- sequestration. Most importantly, we found that it is not possible
mately 0.54 for the same sample. In the case of Sample 23, at the to obtain converging (optimum) results using unipore shale-matrix
same pressure, the ratios are 0.72 and 0.46, respectively. models for the forward simulation; the matrix model used for the
The rapid change in free-gas-CO2 measurements near 950 psia history matching should be a dual-porosity (organic/inorganic)
in Figs. 9 and 10 is caused by dramatic changes in the z factor of system. History-matching results at various confining-stress condi-
the gas near the critical fluid state. tions and using a multiscale transport model are shown in Table 5
The estimated coefficient of isothermal pore-volume-com- for Samples 21 and 23. Note that the time required to reach the new
pressibility values in Table 4 are nearly the same for the two sam- equilibrium is significantly less as the pore pressure is increased.
ples, taking values on the order of from 5.0×10−6 to 7.0×10−6 psi−1. This indicates that the transport paths are sensitive to effective
During our gas-storage measurements and calculations, we found stress. The observation is supported further by the increase in the
that the results would have been less than 5% different when the estimated inorganic permeability and organic diffusivity values.
pore-compressibility effect had been ignored. Thus, the primary We note that permeability estimates are significantly less than
correction necessary for the organic pore volume is the adsorption those obtained using conventional measurements, such as Jones’
effect. This behavior is quite different than in coals where most of method (see also Fig. 11b for details). This is possibly because of
the matrix is organic and the pore-volume-compressibility values the explicit introduction of ultralow-permeability organics into the
could be one or two orders of magnitude larger; hence, its effect description of shale-gas transport in this work.
on the pore volume measurements in coal becomes comparable to It was argued in the Interpretation of Organic Pore Diffusion
the sorption effects (Koleowo 2010). subsection that the pore-diffusion coefficient for a single capil-
Finally, the estimated adsorbed-phase density values in Table lary should typically be two orders of magnitude larger than the
4 for methane and CO2 are in agreement with those reported in the experimentally obtained apparent free-gas transport coefficient
literature (Kurniawan et al. 2006; Ambrose et al. 2010), indicating DK. Here, using Eq. 4, we experimentally predict tortuosity val-
that the gas-storage models are reasonable and that the density for ues within the range of 4.0 to 6.0 for the kerogen. The estimated
the adsorbed-phase is close to the liquid density of the chemical values are consistent with the literature, showing high tortuosity
species used during the experiment. values for microporous materials and suggesting that the mea-
A typical upstream pressure-decline curve of the pulse-decay sured diffusion coefficient indeed represents the coefficient for
gas-permeation experiments is shown in Fig. 11. In this case, pore diffusion of free gas. More interestingly, the coefficient for
methane initially in equilibrium with shale Sample 21 is allowed adsorbed-phase transport DS is predicted to be typically three
to permeate. The curve typically yields two distinct behaviors orders of magnitude larger than that for the free gas. This is a
having completely different characteristic times. At early times, somewhat curious behavior, which suggests that the surface diffu-
the curve shows a rapid decline in pressure (fast transport), and, sion could be the dominant mechanism of transport in the organ-
later, it slows down, gradually decreasing the upstream pressure ics. Typically, the adsorbed-phase concentration and its gradient
(slow transport). These distinct portions of the decline curve in the organics are smaller than those values associated with the
include information related to Darcy flow in the inorganic matrix free gas; however, when multiplied with an extremely large sur-
consisting of large voids, and pore/surface diffusion in the organic face diffusivity, the product (i.e., mass flux of the adsorbed phase)
pore network. In addition, the total time required for the shale/gas becomes a quantity that significantly contributes to the total mass
system to reach the new thermodynamic equilibrium is related flux in the kerogen. The importance of surface diffusion to the
to gas storability for the shale and can be used to obtain basic gas transport in heterogeneous coals and shales has been discussed
information such as organic-/inorganic-pore-volume ratio. Having recently by Fathi (2010) and Fathi and Akkutlu (in press) on the
the pressure-decline curve, we used a simulation-based history- basis of theoretical arguments in the spectral (Fourier-Laplace)
matching algorithm based on the randomized maximum likelihood domain. Here, the experimental results support the observation,
method and estimated the optimized transport properties of the suggesting that the sequestered gases will also be transported as
shale samples. The optimization requires different realizations of part of the adsorbed phase in the organics. Fig. 12 shows that the
the parameter space, typically on the order of hundreds. Details estimated mass flux from surface diffusion of the adsorbed gas
of the automated nonlinear history-matching process are described molecules is comparable to and could be even significantly larger
elsewhere (Oliver et al. 2008). Here, salient results related to the than the mass flux from pore diffusion of the free-gas molecules.

852 December 2011 SPE Journal


TABLE 5—ESTIMATED METHANE TRANSPORT PARAMETERS* FOR ORGANIC-RICH SHALE SAMPLES
USING ISOTHERMAL PULSE-DECAY MEASUREMENTS AT VARYING PORE PRESSURES

Estimated Matrix Parameters for Shale Sample 21


Effective Stress Conditions Pulse
Total
Equilibrium Time To
Confining Pore Reach Jones’ TOC φ Inorganic Kerogen
Pressure Pressure p Equilibrium Permeability (mass/mass) (vol/vol)
(psi) (psi) (psi) (seconds) (nd) % % k (nd) k0 DK (cm2/s) DS (cm2/s)

0 1.73
5 2
5,000 1,623 83 44,485 70.4 3.9 1.33 15.8 53.87 4.45 3.50 10 1.10 10
5 2
5,000 2,588 84 18,063 102.2 3.9 1.24 20.4 64.16 4.39 4.00 10 4.20 10
5 2
5,000 2,974 70 12,060 122.9 3.9 1.22 28.7 67.56 4.04 4.90 10 8.80 10

Estimated Matrix Parameters for Shale Sample 23


Effective Stress Conditions Pulse
Total
Equilibrium Time To
Confining Pore Reach Jones’ TOC φ Inorganic Kerogen
Pressure Pressure p Equilibrium Permeability (mass/mass) (vol/vol)
(psi) (psi) (psi) (seconds) (nd) % % k (nd) k0 DK (cm2/s) DS (cm2/s)

0 2.85
5 3
5,000 1,772 78 72,060 12.65 3 2.4 2.6 43.58 6.14 3.50 10 1.55 10
5 3
5,000 25,008 74 54,480 16.31 3 2.32 3.52 47.09 5.10 5.30 10 3.50 10
5 3
5,000 3,066 73 32,520 17.7 3 2.27 5.24 48.23 4.10 8.20 10 6.35 10
* In the tables, porosity represents the void volume available for free gas only; hence, porosity is corrected for the pore com pressibility and methane-adsorption effects;
TOC is measured independently; k0 represents the ratio of kerogen pore volume to the total pore volume; Dk is the molecular diffusivity of free gas in kerogen.

Further investigation using different shale samples is currently and permeability, and it could be dominated by the adsorbed-
necessary. phase transport taking place in the organics. It is a difficult task
to experimentally differentiate slippage of free-gas molecules
Conclusions and surface-diffusion effects taking place in the phase-transition
In this paper, we have developed methodologies that allow us to region near the organic nanopore walls, although strong depen-
investigate in detail CO2 sequestration in organic-rich gas-shale dence on the adsorbed-phase gradient indicates the significance of
samples from Forth Worth basin. It is found that the pore-volume surface diffusion in the organics. Although the study is important
estimation is a crucial step for the sequestration considerations for our understanding of gas storage and transport in organic-rich
in gas shale. This volume may not be important for the storage shale, gas shales in North America may differ in terms of organic
of free gas, however, because up to 97% of the uptaken gas is content, pore structure of the organic material, and temperature.
stored in adsorbed state inside the organic pores, depending on the Further investigation involving samples from different shale
shale-gas-reservoir pressure and temperature. It is also found that plays is necessary for statistical inference and generalization of
the gas transport takes place in the presence of dynamic porosity the results.

Sample 21 Sample 23
2.5×10–9 2.5×10–9
Free gas diffusive mass flux Free gas diffusive mass flux
JK,x and JS,x(mol-cm/cm3/s)

JK,x and JS,x(mol-cm/cm3/s)

Adsorbed gas diffusive mass flux Adsorbed gas diffusive mass flux
2×10–9 2×10–9
Darcy flux
Darcy flux

1.5×10–9 1.5×10–9
Direction of
gas permeation
1×10–9 1×10–9

5×10–10 5×10–10

0 0
0 1 2 3 4 5 0 1 2 3
Distance, cm Distance, cm
(a) (b)

Fig. 12—Predicted mass fluxes across the shale Sample 21 at initial equilibrium pore pressure of 2,974 psia and Sample 23 at
3,066 psia during the pressure-pulse-decay measurements. The mass fluxes are computed using forward simulation with the
optimized transport parameters shown in Table 5. Only the flux values at the halftime to reach the new equilibrium pressure
during the pulse-decay measurement are shown.

December 2011 SPE Journal 853


Nomenclature Gas Conference, Pittsburgh, Pennsylvania, 23–25 February. doi:
As = surface area, nm2 10.2118/131772-MS.
C = moles of concentration Anderson, J.R. and Pratt, K.C. 1985. Introduction to Characterization and
Cp = the pore-volume compressibility, psi−1 Testing of Catalysts. Sydney, Australia: Academic Press.
Cµ = adsorbed-phase amount in moles per solid organic Bear, J. and Bachmat, Y. 1991. Introduction to Modeling of Transport
volume, mol/cm3 Phenomena in Porous Media, paperback edition. Dordrecht, The
DK = pore-diffusion coefficient, cm2/s Netherlands: Theory and Applications of Transport in Porous Media,
DK, capillary = molecular-diffusion coefficient for gas transport in the Kluwer Academic Publishers.
capillary, cm2/s Brace, W.F., Walsh, J.B., and Frangos, W.T. 1968. Permeability of Gran-
DS = porosity- and tortuosity-corrected surface diffusivity, ite Under High Pressure. J. Geophys. Res. 73 (6): 2225–2236. doi:
cm2/s 10.1029/JB073i006p02225.
JK,x = pore-diffusion mass flux, mol·cm/cm3/s Busch, A., Alles, S., Gensterblum, Y., Prinz, D., Dewhurst, D.N., Raven,
JS,x = surface-diffusion mass flux, mol·cm/cm3/s M.D., Stanjek, H., and Krooss, B.M. 2008. Carbon Dioxide Storage
n = number of moles, mol Potential of Shales. International Journal of Greenhouse Gas Control
n1 = moles of gas in V1, mol 2 (3): 297–308. doi: 10.1016/j.ijggc.2008.03.003.
n1f = moles of gas in V1 at Pf , mol Fathi, E. 2010. Fluid Flow, Transport and Reaction in Heterogeneous
n2 = moles of gas in V2, mol Porous Media. PhD thesis, University of Oklahoma, Norman, Okla-
n2f = moles of gas in V2 at Pf , mol homa.
nsp = moles of gas stored in the sample, mol Fathi, E. and Akkutlu, I.Y. 2009. Matrix Heterogeneity Effects on Gas
NKn = Knudsen number Transport and Adsorption in Coalbed and Shale Gas Reservoirs.
P = pore pressure Transport in Porous Media 80 (2): 281–304. doi: 10.1007/s11242-
P1 = equilibrium pressure in V1, psia 009-9359-4.
P2 = initial pressure in V3, psia Fathi, E. and Akkutlu, I.Y. In press. Mass Transport of Adsorbed-phase
Pf = new equilibrium pressure in dead volume and sample in Stochastic Porous Medium with Fluctuating Porosity Field and
volume, psia Nonlinear Gas Adsorption Kinetics. Transport in Porous Media (sub-
PL = Langmuir pressure, psia mitted 2010).
Ps = pore pressure in the sample, psia Finsterle, S. and Persoff, P. 1997. Determining Permeability of Tight
Sa = adsorbed-gas storage, mol Rock Samples Using Inverse Modeling. Water Resour. Res. 33 (8):
Samax = total moles of adsorbed gas at infinite pore pressure, 1803–1811.
mol Jones, S.C. 1997. A Technique for Faster Pulse-Decay Permeability Mea-
Sw = water saturation, % surements in Tight Rocks. SPE Form Eval 12 (1): 19–26. SPE- 28450-
V1 = reservoir volume, cm3 PA. doi: 10.2118/28450-PA.
V2 = dead volume, cm3 Koleowo, O.B. 2010. Gas Storage Capacity and Transport in Coals. MS
Va = adsorbed-gas volume, cm3 thesis, University of Oklahoma, Norman, Oklahoma.
Vamax = maximum adsorbed gas, cm3 Kurniawan, Y., Bhatia, S.K., and Rudolph, V. 2006. Simulation of Binary
VL = Langmuir volume, scf/ton Mixture Adsorption of Methane and CO2 at Supercritical Conditions in
Vp = sample pore volume, cm3 Carbons. AIChE Journal 52 (3): 957–967. doi: 10.1002/aic.10687.
Vp0 = pore volume at zero pore pressure, cm3 Lee, B. and Tan, T. 1987. Application of Multiple Porosity/Permeability
Vp2 = pore volume at P2, cm3 Simulator in Fractured Reservoir Simulation. Paper SPE 16009 pre-
Vpf = pore volume at Pf, cm3 sented at the SPE Symposium on Reservoir Simulation, San Antonio,
z1 = correction factor at P1 Texas, USA, 1–4 February. doi: 10.2118/16009-MS.
z2 = correction factor at P2 Loucks, R.G., Reed, R.M., Ruppel, S.C., and Jarvie, D.M. 2009. Morphol-
zf = correction factor at Pf ogy, Genesis, and Distribution of Nanometer-Scale Pores in Siliceous
zs = correction factor at Ps Mudstones of the Mississippian Barnett Shale. Journal of Sedimentary
n1 = change of moles of gas in V1, mol Research 79 (12): 848–861. doi: 10.2110/jsr.2009.092.
n2 = change of moles of gas in V2, mol Ning, X. 1992. The Measurement of Matrix and Fracture Properties
ns = change of moles of gas in sample, mol in Naturally Fractured Low Permeability Cores using a Pressure
nsa = change of moles of adsorbed gas in sample, mol Pulse Method. PhD thesis, Texas A&M University, College Station,
nsp = change of moles of gas stored in sample, mol Texas.
Vpa = change of volume of adsorbed gas in sample, cm3 Ozdemir, E., Morsi, B.I., and Schroeder, K. 2004. CO2 Adsorption Capac-
Vpp = change of pore volume at the pore pressure in ity of Argonne Premium Coals. Fuel 83 (7–8): 1085–1094.
sample, cm3 Oliver, D.S., Reynolds, A.C., and Liu, N. 2008. Inverse Theory for Petro-
k0 = organic pore volume to total pore volume, ratio leum Reservoir Characterization and History Matching. Cambridge
TOC = kerogen to total volume ratio University Press.
 = bulk density of gas, g/cm3 Roy, S., Raju, R., Chuang, H.F., Cruden, B.A., and Meyyappan, M. 2003.
ads = molar density of adsorbed gas, mole/cm3 Modeling Gas Flow Through Microchannels and Nanopores. J. Appl.
 = tortuosity Phys. 93 (8): 4870–4879. doi: 10.1063/1.1559936.
 = porosity Schettler, P.D., Parmely, C.R., Juniata, C., and Lee, W.J. 1989. Gas Stor-
age and Transport in Devonian Shales. SPE Form Eval 4 (3): 371–376;
Trans., AIME, 287. SPE-17070-PA. doi: 10.2118/17070-PA.
Acknowledgments
Sondergeld, C.H., Ambrose, R.J., Rai, C.S., and Moncrieff, J. 2010. Micro-
We recognize the work and contributions of Gary D. Stowe in the Structural Studies of Gas Shales. Paper SPE 131771 presented at the
IC3 laboratory at The University of Oklahoma. We thank Mery SPE Unconventional Gas Conference, Pittsburg, Pennsylvania, USA,
Diaz-Campos for her contribution to the generation of Fig. 2. 23–25 February 2010. doi: 10.2118/131771-MS.
For insightful discussions, we thank Chandra S. Rai and Carl H. Wang, F.P. and Reed, R.M. 2009. Pore Networks and Fluid Flow in
Sondergeld in shale petrophysics and Dean S. Oliver in nonlinear Gas Shales. Paper SPE 124253 presented at the SPE Annual Tech-
optimization. nical Conference and Exhibition, New Orleans, 4–7 October. doi:
10.2118/124253-MS.
References Wei, X.R., Weng, G.X., Massarotto, P., Rudolph, V., and Golding, S.D.
Ambrose, R.J., Hartman, R.C., Diaz-Campos, M., Akkutlu, I.Y., and Son- 2007. Modeling Gas Displacement Kinetics in Coal With Maxwell-
dergeld, C.H. 2010. New Pore-scale Considerations in Shale Gas in Place Stefan Diffusion Theory. AIChE Journal 53 (12): 3241–3252. doi:
Calculations. Paper SPE 131772 presented at the SPE Unconventional 10.1002/aic.11314.

854 December 2011 SPE Journal


Seung Mo Kang is an MS student in petroleum engineer- of gas-storage mechanisms in unconventional gas resources,
ing at The University of Oklahoma. He holds a BS degree in and estimation of gas in place. I. Yucel Akkutlu is professor
mechanical engineering from Hanyang University in South of petroleum engineering at The University of Oklahoma. He
Korea. His current research interest is quantification of gas- holds MS and PhD degrees from the University of Southern
storage mechanisms in shale gas reservoirs. Ebrahim Fathi is a California. His work finds applications in reservoir engineering,
research associate at The University of Oklahoma. Fathi holds particularly in the areas of unconventional gas recovery and
BS and MS degrees in exploration of mining and petroleum improved and enhanced oil recovery. His current research
engineering from Tehran University. He holds a PhD degree in deals with the thermodynamics of fluids in nanoporous materi-
petroleum engineering from The University of Oklahoma. His als and with the scaling up of coupled transport and reaction
research interests are up-scaling of the fluid flow, transport processes in low-permeability geological formations exhibiting
and reaction processes in heterogeneous porous media, gas multiscale pore structures. Richard F. Sigal joined The University
transport and storage in shale gas and coalbed methane res- of Oklahoma in 2004 as Unocal Centennial Professor with a
ervoirs, and simulation of flow dynamics in naturally occurring joint appointment in petroleum engineering and geology and
nanoporous materials. Raymond J. Ambrose is a PhD student geophysics. His primary research activity since then has been
in petroleum engineering at The University of Oklahoma and on tight gas, shale gas, and other unconventional gas reser-
director of reservoir engineering for Reliance Holding USA. He voirs. Before joining The University of Oklahoma, he worked in
holds a BS degree in chemical engineering and an MS degree the oil and gas industry for 26 years on reservoir characteriza-
in petroleum engineering, both from the University of Southern tion, core measurements, and petrophysics. Among his areas
California. His current research interests are analytical solu- of special expertise are nuclear magnetic resonance, mercury
tions for gas shale productivity, SEM imaging and pore struc- capillary pressure measurements, and the stress dependence
ture characterization for organic-rich gas shale, identification of permeability.

December 2011 SPE Journal 855


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 337 Total Pages: 10

Simultaneous Multifracture Treatments:


Fully Coupled Fluid Flow and Fracture
Mechanics for Horizontal Wells
Kan Wu, SPE, and Jon E. Olson, SPE, University of Texas at Austin

Summary Layne (1991), Mack et al. (1992), and Elbel et al. (1992), in which
Successfully creating multiple hydraulic fractures in horizontal fracture propagation was modeled according to the Geerstma and
wells is critical for unconventional gas production economically. DeKlerk geometry, the Perkins and Kern geometry, and a pseudo-
Optimizing the stimulation of these wells will require models that 3D geometry, respectively, with fluid flow in multiple fractures and
can account for the simultaneous propagation of multiple, poten- the wellbore, but with no mechanical interaction between multiple
tially nonplanar, fractures. fractures nor independent pressure drop in each fracture. Olson
In this paper, a novel fracture-propagation model (FPM) is (2008) used a pseudo-3D approach for vertical but laterally nonpla-
described that can simulate multiple-hydraulic-fracture propagation nar fractures, assuming zero-viscosity injection fluid to reduce
from a horizontal wellbore. The model couples fracture deformation computation time for multifracture problems. Roussel and Sharma
with fluid flow in the fractures and the horizontal wellbore. The dis- (2011) used a 3D finite-difference numerical model to calculate
placement discontinuity method (DDM) is used to represent the stress distributions around uniform-pressure fractures, accounting
mechanics of the fractures and their opening, including interaction for propagation by assuming fractures would follow principal-stress
effects between closely spaced fractures. Fluid flow in the fractures trajectories. Weng et al. (2011) and Wu et al. (2012) describe a
is determined by the lubrication theory. Frictional pressure drop in model for simulating multiple hydraulic fractures in naturally frac-
the wellbore and perforation zones is taken into account by applying tured reservoirs combining standard pseudo-3D equations for frac-
Kirchoff’s first and second laws. The fluid-flow rates and pressure ture-width calculations coupled with the displacement discontinuity
compatibility are maintained between the wellbore and the multiple approach of Olson (2004) to account for stress-interaction effects
fractures with Newton’s numerical method. The model generates between fractures. Meyer and Bazan (2011) presented a discrete-
physically realistic multiple-fracture geometries and nonplanar-frac- fracture-network model to simulate planar-fracture growth in the
ture trajectories that are consistent with physical-laboratory results three principal planes, with the Olson (2004) analytical solution to
and inferences drawn from microseismic diagnostic interpretations. describe the interaction between induced fractures. Wong and Xu
One can use the simulation results of the FPM for sensitivity (2013) modeled multiple-fracture propagation in horizontal wells
analysis of in-situ and fracture treatment parameters for shale-gas with mechanical interaction through the boundary integral formula-
stimulation design. They provide a physics-based complex frac- tion. Castonguay et al. (2013) implemented the weakly singular,
ture network that one can import into reservoir-simulation models symmetric Galerkin boundary-element method to calculate fracture
for production analysis. Furthermore, the results from the model geometry, in which flow in the fracture is modeled as power-law
can highlight conditions under which restricted width occurs that fluid flow in the arbitrary curved channels. Nagel et al. (2011) pro-
could lead to proppant screenout. posed a discrete-element model, in which the discrete-fracture net-
work consists of planar, finite fractures in three dimensions that
bound intervening rigid or deformable reservoir blocks. Dahi Tale-
Introduction ghani and Olson (2011) presented a coupled fluid-flow and rock-
Multiple hydraulic fracturing in combination with horizontal dril- mechanics model that was based on the extended finite-element
ling significantly improves well productivity in shale reservoirs. method to simulate the fracture geometry of plane-strain problems.
Limited-entry fracturing techniques are widely used to create Keshavarzi and Mohammadi (2012) also applied the extended fi-
simultaneously propagating multiple fractures (Daneshy 2011). In a nite-element method to analyze the interaction between hydraulic
single fracturing stage, there can be three to six perforation clusters. and natural fractures under plane-strain and quasistatic conditions.
In comparison with modeling single fracturing propagation, multi- The model presented in this paper follows the single nonplanar-
ple-fracture-propagation modeling needs to consider mechanical fracture model of Olson and Wu (2012). The model incorporates
fracture interaction, which is a dynamic process of subsurface geo- fracture deformation and non-Newtonian fluid flow in the wellbore
mechanical stress changes induced by fracture growth. The signifi- and fractures. Fracture widths and stress-intensity factors are calcu-
cant consequences of multifracture treatments are stress-shadow lated with the modified, pseudo-3D DDM approach of Olson
effects that lead to complex fracture geometry and fracture-width (2004), accounting for all mechanical interactions between every
restriction. Fracture geometry can be further complicated by pre- fracture element at every timestep. The flow rate in each fracture
existing natural fractures in shale-gas reservoirs (Gale et al. 2007; satisfies mass balance and pressure equilibrium between the hori-
Olson 2008; Chong et al. 2010), which is not the focus of this work zontal wellbore and hydraulic fractures. The Newton-Raphson nu-
but which will be discussed in a future publication. To optimize merical method is applied to solve nonlinear fluid-flow equations.
multiple-fracture treatments, characterization of the fracture geom- The Picard iterative method is used to fully couple fracture
etry is very necessary. However, the complex-fracture geometry mechanics and fluid flow. We validate our model against two ana-
cannot be measured directly and completely by any fracture diag- lytical solutions and published numerical results. A field case study
nostic techniques. Therefore, modeling remains an important tool is performed to predict multiple-fracture geometry and to calculate
for engineers to predict the fracture geometry. stress distribution with the purpose of demonstrating the feasibility
Recent efforts to model multifracture treatments have taken var- of the model.
ious forms. Multicrack models were described by Siriwardane and
Fracture Mechanics
Copyright V
C 2015 Society of Petroleum Engineers
Fracture mechanics determines fracture displacements and propa-
This paper (SPE 167626) was accepted for presentation at the SPE Annual Technical gation direction. The 2D DDM (Crouch 1976) with a 3D correc-
Conference and Exhibition, New Orleans, 30 September–2 October 2013, and revised for
publication. Original manuscript received for review 1 November 2013. Revised manuscript tion factor (Olson 2004) is used to calculate normal and shear
received for review 8 March 2014. Paper peer approved 17 April 2014. displacements considering mechanical interaction. The fracture-

April 2015 SPE Journal 337

ID: jaganm Time: 11:58 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 338 Total Pages: 10

y y

Dx Dy

x z
σzz
2a

σxx y

h
+
S x

S

x
Fig. 2—Sketch of a uniformly loaded, isolated, vertical fracture
Fig. 1—Illustration of a 2D boundary with crack-like geometry. with a finite height and an infinite length.

propagation direction is determined by the maximum-circumfer- Because hydraulic fractures are best modeled accounting for
ential-stress theory (Erdogan and Sih 1963), a criterion based on their 3D geometry (finite height and length), we use a 3D correc-
the stress-intensity factors at the crack tips. tion factor Gij derived by Olson (2004),

DDM. The 2D DDM, a special boundary-element method, was dijb


Gij ¼ 1  h ib=2 ; . . . . . . . . . . . . . . . . . . . ð3Þ
developed by Crouch (1976) and designed for handling problems
with crack-like geometries. Stresses at a point n within an elastic dij2 þ ðh=aÞ2
region are integrals of displacements Dui over the boundary S6 of
the region (Fig. 1), where dij is the distance between the centers of fracture elements i
ð and j, h is the fracture height, and a and b are empirically deter-
mined constants found by comparing displacements and stresses
rjk ðnÞ ¼ Eijk ðn; gÞDui ðgÞdSðgÞ
S ; . . . . . . . . . . . . . ð1Þ with a true 3D fracture solution (Shou 1993). This correction fac-
þ
Dui ¼ ui  u i
tor was modeled after the analytical plane-strain equation for the
normal stress, rxx, acting perpendicular to a uniformly loaded, iso-
where the quantity Eijk(n,g) is the tensor field that represents the lated, vertical crack of finite height and infinite length (Pollard
influences of a concentrated force at point n on displacements at and Segall 1987). The use of a ¼ 2, b ¼ 3, and dij ¼ x in Eq. 3
point g. Because analytical solutions are very hard to obtain for returns the dimensional factor from the following equation for
Eq. 1, a numerical solution was applied to deal with this problem. fracture-induced normal stress rxx along the x-axis (Fig. 2),
The boundary S6 is divided into N planar elements. An analytical 8 9
solution of a constant displacement-discontinuity element in an >
< 3 >
=
infinite elastic solid is obtained with the Green’s-function ap- jxj
rxx ðxÞ ¼ DrI 1  h i : . . . . . . . . . . . . ð4Þ
proach that is based on the solution of Kelvin’s problem. On the >
:
3=2 >
;
x2 þ ðh=2Þ2
basis of the analytical solution, one can calculate stresses at arbi-
trary points with the principle of superposition by linearly com-
bining the contributions of the displacement discontinuities of all Maximum-Circumferential-Stress Criterion. The opening
elements. mode and sliding-shear mode stress-intensity factors, KI and KII,
For fracture problems, displacement or stress-boundary condi- are functions of normal- and shear-displacement discontinuities at
tions are generally defined on the crack surfaces along with any the crack tip element, Dn and Ds, respectively, because (Olson
remote-boundary conditions. In addition, the simultaneous solu- 1990, 2007)
tion of the system of equations accounts for stresses caused by the pffiffiffi
mechanical interaction with other nearby elements (Crouch and 0:806E p
Starfield 1983). Thus, because of the normal- (rin ) and shear- (ris ) KI ¼ pffiffiffiffiffi Dn
4ð1   2 Þ 2a
stress boundary conditions on each element i determined with the and pffiffiffi . . . . . . . . . . . . . . . . . . . . . ð5Þ
fluid-flow solution, we can determine the normal- and shear-dis- 0:806E p
placement discontinuities, Dnj and Dsj , respectively, at each ele- KII ¼ pffiffiffiffiffi Ds :
4ð1   2 Þ 2a
ment j by solving a linear system of equations given by
These expressions were shown to be accurate for stress-inten-
X
N X
N
rin ¼ Gij Cijns Dsj þ Gij Cijnn Dnj sity-factor determination in fully 3D displacement-discontinuity
j1 j1
models as well (Sheibani and Olson 2013). Knowing the stress-in-
: . . . . . . . . . . . . . . . ð2Þ tensity factors, one can determine the fracture-propagation direc-
X
N X
N
ris ¼ Gij Cijss Dsj þ Gij Cijsn Dnj tion h by the maximum-circumferential-stress theory (Erdogan
j1 j1 and Sih 1963) as
ij ij
C and G are factors that account for the elastic mechanical KI sinh þ KII ð3cosh  1Þ ¼ 0; . . . . . . . . . . . . . . . . . . . ð6Þ
interaction between elements. Cijns is the plane-strain, elastic-influ-
ence coefficient matrix representing the normal stress at element i where h is the counterclockwise positive angle from the in-plane
induced by a shear-displacement discontinuity at element j, and propagation of the fracture ahead of the tip. For pure opening
Cijnn gives the normal stress at element i because of an opening dis- mode (KI > 0 and KII ¼ 0), we would predict planar-fracture prop-
placement discontinuity at element j. We can attribute analogous agation along h ¼ 0 . For pure sliding mode (KI ¼ 0 and KII = 0),
meanings to Cijss and Cijsn . The detailed derivation of C is given by the strongest possible kink in fracture direction occurs along
Crouch and Starfield (1983). h ¼ 670.5 for positive or negative KII, respectively.

338 April 2015 SPE Journal

ID: jaganm Time: 11:58 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 339 Total Pages: 10

Fluid Flow pcf,i is pressure loss in the horizontal wellbore given by (Valko
Multiple-fracture propagation is driven by an incompressible and and Economides 1995)
non-Newtonian fluid. Fluid flow consists of two parts—fluid flow X
i
0
in the fracture and fluid flow in the horizontal wellbore between pcf ;i ¼ Ccf ðxj  xj1 ÞQwj n ;
injection points. j¼1
2ðX
j1Þ
Fluid Flow in a Single Fracture. Fluid flow through a rock frac- Qwj ¼ QT  Qk ;
ture is modeled by the Navier-Stokes equations of fluid mechanics k¼1 : . . . . . . . . . .ð11Þ
(Schlichting 1968). When assuming a uniform pressure gradient Qwj ¼ QT at j ¼ 1;
in the channel between two parallel and smooth fracture surfaces,
one can simplify the Navier-Stokes equations (Valko and Econo- and  n0
mides 1995): 0 0 1 þ 3n0 0
Ccf ¼ 23n þ2 pn k0 Dð3n þ1Þ :
 0 n0
@p 0 1 þ 2n0 n n0 ð2n0 þ1Þ n0
¼ 2n þ1 k0 h w Q : . . . . . . . . . . ð7Þ
@s n0 The pressure drop between the wellbore intervals is a function of
distance between two neighboring fractures and flow rate Qwj in the
From Eq. 7, we can obtain the relationship between pressure drop wellbore intervals. The longer the distance between two neighboring
and fluid properties, fracture geometry, and flow rate. Pressure fractures, the larger the pressure drop. This is a rough approach to esti-
drop increases with increasing flow rate Q and decreases with mate frictional pressure drop in the wellbore without considering
increasing fracture width w. In our model, to enhance conver- flowing conditions. Siriwardane and Layne (1991) gave a more-accu-
gence stability with little cost in accuracy, the flow-rate gradient rate approach to calculate the pressure loss in the wellbore by taking
along the fracture is ignored. Spence and Sharp (1985) and Zim- into account friction factors under turbulent and laminar conditions.
merman et al. (1991) discussed the validation of the lubrication
theory for modeling fluid flow inside the fractures. According to Material Balance. One must satisfy material balance in the
their analysis, fracture aperture cannot change too abruptly and whole fluid-flow system. The total volume of fluid pumped during
should be small in comparison to the other dimensions of the frac- injection time is equal to the fluid volume preserved in the multi-
tures, and fluid flow should be laminar. ple fractures and the volume leaking into the surrounding porous
medium. The global volume balance is
Fluid Flow in Horizontal Wellbore. A numerical model for
multilayer-fracturing treatments in vertical wells was presented ðt i ðtÞ
Lð i ðtÞ ð
Lð t
X
N X
N
by Siriwardane and Layne (1991), Mack et al. (1992), and Elbel QT ðtÞdt ¼ hwds þ qL ðs; tÞdtds: . . . . ð12Þ
et al. (1992). The model described the conservation of volume 1 1
0 0 0 0
and the continuity of pressure by use of a set of constraints, such
as hydrostatic-pressure change, perforation frictional pressure The leakoff volume is described by the Carter leakoff model
drop, and frictional pressure loss in each casing interval. Fluid (Carter 1957), which is expressed as
flow in the horizontal wellbore is similar to fluid flow in the verti-
cal wellbore with multiple layers. The only difference is that we 2hCL
qL ðs; tÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : . . . . . . . . . . . . . . . . . . . . . . . ð13Þ
assume no hydrostatic-pressure change in the horizontal wellbore. t  sðsÞ
Fluid flow in the wellbore is analogous to the flow of electric
current through an electrical circuit network, by applying Kirch- The model assumes that leakoff is 1D in the direction perpen-
off’s first and second laws to obtain the flow-rate distribution of dicular to the fracture plane.
multiple fractures (Siriwardane and Layne1991; Mack et al. 1992;
Elbel et al. 1992). Our model assumes that multiple-transverse Iterative Algorithm
fractures are started perpendicular to the horizontal wellbore at
The fracture mechanics and fluid flow are coupled numerically
perforation-cluster locations. The total injection rate QT is divided
through an iterative algorithm. The elastic DDM solution deter-
into injection rates of each fracture Qi. The flow-rate distribution
is a dynamic process, and the diversion of the fluid into various mines the displacement discontinuities (opening and shearing) for
perforation clusters is a function of wellbore friction, perforation each fracture element because of its internal pressure and resolved
friction, and the fluid pressure within the fractures (Fig. 3). external stresses, whereas the internal-pressure distribution is
By ignoring wellbore-storage effects, total injection rate related to fracture opening by the fluid flow. The mechanical solu-
tion is a linear set of equations, whereas the fluid flow is nonlin-
should be the sum of the injection rates of all fractures,
ear. Fluid flow is discretized by the finite-difference method on a
X
N fixed grid of boundary elements. Newton’s numerical method is
QT ¼ Qi : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð8Þ used to solve the system of nonlinear equations. Picard coupling
i¼1 (Li 1991; Adachi et al. 2007) is used to couple the fracture
mechanics and fluid flow at every updated timestep. If the conver-
Kirchoff’s second law describes the continuity of pressure, in gence criterion is not satisfied, the net pressure is modified at each
which the pressure in the wellbore heel is equal to the sum of iteration step as
wellbore frictional pressure drop, perforation frictional pressure
drop, and pressure in the first element of the fracture as (Elbel pkþ1 ¼ ð1  ar Þpk1 þ ar pk : . . . . . . . . . . . . . . . . . . . ð14Þ
et al. 1992)
X
i Displacement discontinuities are recalculated with the modi-
P0 ¼ pw;i þ ppf ;i þ pcf ; j : . . . . . . . . . . . . . . . . . . . ð9Þ fied pressure from the previous iteration step. The convergence
j¼1 criterion is checked again after computing the new net pressure
on the basis of the displacements. ar is a retardation coefficient
pw,i is pressure in the first element of fracture, which is calculated ranging from 0 to 0.5 that dampens variations to avoid stress
by Eq. 7; ppf,i is the perforation frictional pressure loss that is pro- jumps and extreme changes in pressure. If the convergence crite-
portional to the square of flow rate, estimated by (Elbel et al. rion cannot be satisfied, decreasing ar is an approach to improve
1992) convergence. Smaller ar can cause convergence to be too slow,
qs whereas a large ar can result in diverging oscillations. We found
ppf ;i ¼ Qi 2 : . . . . . . . . . . . . . . . . . . . ð10Þ that ar ¼ 0.2 provides a reasonably stable convergence. The flow
0:323qw n2p;i dp;i
4
chart in Fig. 4 describes the structure of the program.

April 2015 SPE Journal 339

ID: jaganm Time: 11:58 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 340 Total Pages: 10

Frac. 1 Frac. 2 Frac. n

Q1 Pw,1 Q3 Pw,3 Q2n-1 Pw,2n-1

QT,P0 Ppf,1 QT-Q1-Q2 Ppf,3 QT – (Q1+...+ Q2n-2) Ppf,2n-1

Ppf,2 ΔPcf,2 Ppf,4 ΔPcf,n Ppf,2n


ΔPcf,1

Q2 Pw,2 Q4 Pw,4 Q2n Pw,2n

Fig. 3—Illustration of flow-rate partition and pressure drop.

Model Validation pared with the PKN model. At the earlier time of propagation,
Comparison With Analytical Solutions. Our FPM extends from KGD gives more-accurate results than PKN because fracture length
a nonplanar single-fracture model to a multiple-fracture model. is shorter than fracture height. As fracture length grows, KGD grad-
The single-fracture, nonplanar model developed by Olson and Wu ually loses its accuracy, and PKN can give more-accurate results.
(2012) was validated against planar pseudo-3D models from the The other input parameters are listed in Table 2.
Gas Research Institute Staged Field Experiment No. 3 case (War- A comparison of results made for fracture length (Fig. 5), frac-
pinski et al. 1994). In this paper, we compare our model to the ture width at the wellbore (Fig. 6), and net injection pressure
end-member analytical solutions of Khristianovich-Geertsma- (Fig. 7) is presented. As shown, our model can give a very good
Deklerk (KGD) and Perkins-Kern-Nordgren (PKN) (Table 1). agreement with the KGD in the earlier time and tends to match
The KGD model presumes plane strain in the horizontal plane and PKN analytical solutions as the fracture grows longer, which
a rectangular vertical cross section (no width variation over the lends credence to our model results and analysis.
height). The PKN model assumes plane strain in a vertical plane
normal to the fracture-propagation direction, and the vertical cross Comparison With Unconventional Fracture Model. For multi-
section of the fracture is elliptical in shape. To compare with the ple-fracture propagation, fracture interaction plays an important
KGD and PKN solutions, we simulate a case with a fracture height role in fracture geometry. For nonplanar propagation validation,
of 200 ft. The fracture geometry calculated by KGD is much fatter we refer for comparison to the nonplanar model of Weng et al.
and shorter with very low net pressure that climbs with time, com- (2011) and Wu et al. (2012) for the configurations of a single

Initialization

A new timestep

Element length Add a new element Propagation direction

Displacement
Fracture mechanics
Stress shadow
Pressure
Fluid flow
Flow rate

No Converge

Yes

Mass balance

Check treatment time Continue

Finished

Postprocessing

Fig. 4—Flow chart of the numerical scheme in the model.

340 April 2015 SPE Journal

ID: jaganm Time: 11:58 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 341 Total Pages: 10

TABLE 1—ANALYTICAL SOLUTION OF KGD AND PKN (VALKO AND ECONOMIDES 1995)

Fracture Wellbore Injection


Half-Length Width Pressure
!1=6 !1=6
KGD
i3 E0 li3 Shmin þ 1:09ðE0 2 lÞ1=3 t1=3
0:539 t2=3 2:36 t1=3
lh3f E0 h3f
!1=5  2 1=5 !1=5
i3 E0 4=5 li E0 4 li2
PKN 0:524 t 3:04 0 t1=5 Shmin þ 1:52 t1=5
lh4f E hf h6f

1200
TABLE 2—INPUT PARAMETERS
FPM KGD PKN
1000

Fracture half-length, ft
l 100 cp
Injection Rate 20 bbl/min 800
Young’s Modulus 6,530,000 psi
Poisson’s Ratio 0.2 600
Height 200 ft
400
Minimum Horizontal Stress 4,450 psi
200

0.4 0
Average width at the wellbore, in.

FPM KGD PKN 0 10 20 30 40 50 60 70


0.35
Injection time, min
0.3

0.25 Fig. 5—Comparison of fracture half-length vs. injection time


between the KGD and PKN models and our model.
0.2

0.15
1200
0.1 FPM KGD PKN
Net injection pressure, psi

0.05 900
0
0 10 20 30 40 50 60 70
Injection time, min 600

Fig. 6—Comparison of fracture width at the wellbore between


the KGD and PKN models and our model. 300

horizontal wellbore with two initial fractures (Fig. 8a) and two
horizontal wellbores with one fracture each (Fig. 8b). We simu- 0
0 10 20 30 40 50 60 70
lated these two configurations under isotropic and anisotropic in-
situ stress conditions (see Table 3 for input parameters). The dif- Injection time, min
ferential stress for the anisotropic-stress reservoir is 130 psi, with
the maximum horizontal stress oriented in the y-direction. Fig. 7—Comparison of fracture pressure at the wellbore
between the KGD and PKN models and our model.
Two fractures, started parallel with each other in a horizontal
wellbore for isotropic and anisotropic far-field stresses, are illus- trated in Fig. 9. The two fractures propagate away from each other
because of the mechanical fracture interaction (i.e., stress-shadow
effects). An opening fracture exerts additional stresses on sur-
rounding rock and adjacent fractures, which can result in a local
change in the direction of maximum horizontal stress and a devia-
tion of the fracture path from a planar geometry. In addition, the
fractures have larger curvature under the isotropic-stress condition
than that under the anisotropic condition. This is because the ani-
sotropic far-field stresses tend to suppress fracture curving (Olson
and Pollard 1989). The stress shadow and the anisotropic far-field
stresses compete with each other. To quantify the competition,
Olson and Dahi Taleghani (2009) defined a parameter called the
relative net pressure Rn,
ðPfrac  Shmin Þ
Rn ¼ ; . . . . . . . . . . . . . . . . . . . . . . . ð15Þ
ðShmax  Shmin Þ

(a) (b) where the numerator is the net pressure and the denominator is hori-
zontal differential stress. Stress-shadow effects strongly depend on
Fig. 8—Two initiation fractures in a horizontal wellbore (a) and the net pressure within the fractures (Pollard and Segall 1987; War-
two horizontal wellbores (b); blue lines represent horizontal pinski and Branagan 1989). Larger net pressure will lead to stronger
wellbores, and red lines represent initial fractures. stress-shadow effects. When Rn is much greater than unity, the net

April 2015 SPE Journal 341

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 342 Total Pages: 10

in a horizontal wellbore. We explore the details of simultaneous


TABLE 3—INPUT PARAMETERS (AFTER WU ET AL. 2012) fracture propagation from multiple clusters by studying a case of
six initial fractures investigating propagation geometry, pressure
Injection Rate 40 bbl/min distribution, and stress-shadow effects. The input parameters
Stress Anisotropy 130 psi (Table 4) are based on the published data for the Barnett shale.
Poisson’s Ratio 0.35 Fig. 11 displays the fracture trajectory and fracture width at the
Young’s Modulus 4.35106 psi end of 12 minutes of injection time. This fracture geometry is
Fluid Viscosity 1 cp dominated by two factors—distance from injection point and me-
Fracture Height 394 ft chanical interaction between fractures. To clearly illustrate fracture
Minimum Horizontal Stress 6,773 psi aperture distribution for each fracture, we enlarge the aperture 500
Maximum Horizontal Stress 6,903 psi
times in Fig. 11b. The injection point is at the heel of the horizontal
wellbore (–200, 0). Fractures close to the heel obtain more fluid
Distance Between Initiation Points 33 ft
because of viscous dissipative flow through the wellbore from ear-
lier fractures. Fractures obtaining more fluid have greater length
pressure is much greater than differential stress, which suggests that and aperture. Fractures 1 and 6 are under the same stress-shadow
stress-shadow effects will dominate the fracture geometry. However, position because they are on the outside of the array, but Fracture 1
if the Rn is much less than unity, the fractures will propagate along a is closer to the injection point. Therefore, Fracture 1 has higher net
straight line under the domination of differential stress. However, pressure and grows longer with larger width compared with Frac-
this parameter only indicates whether a nonplanar fracture-propaga- ture 6. All of the interior fractures (Fractures 2, 3, 4, and 5) have
tion path is possible, which means that strong stress-shadow effects similar trends. The space of 50 ft puts adjacent fractures within the
still affect fracture width even if the fractures are perfectly straight in region of stress-shadow influence of their neighbor, resulting in a
a strongly anisotropic stress field. As seen from Fig. 9, we found that strong mechanical interaction that affects propagation path and
these two models give a very similar fracture-propagation path with width. Fractures 2 and 5 are close to the exterior Fractures 1 and 6,
side-by-side configuration. respectively, and are strongly affected by induced stresses of the
Fig. 10 shows two fractures in an en-echelon arrangement fractures, slowing the growth rate of Fractures 2 and 5 and causing
from two adjacent horizontal wellbores under isotropic and aniso- them to deviate from their original propagation path (Fig. 12). Frac-
tropic far-field stress conditions. The two fractures grow toward tures 3 and 4 are less affected by the two exterior fractures. In gen-
each other, because of induced shear stress at the fracture tips eral, the interior fractures are much narrower and shorter compared
caused by mechanical interaction. In the anisotropic far-field case, with the exterior fractures (Fig. 12); this implies a greater flow re-
there is a reduced curvature of fracture paths compared with the sistance for these fractures that diverts fluid elsewhere. The net
isotropic case. In summary, our model compares favorably with pressures for the interior fractures are almost the same as that of the
the nonplanar results generated by Wu et al. (2012). exterior fractures (they all have to have roughly the same net pres-
sure at the injection point because of the low pressure drop in the
wellbore) (Fig. 13). For that to be true, because the widths are less,
A Case Study for Six Fractures Propagating the flow rate has to be less into those fractures. Consequently, inte-
Simultaneously rior fractures have less ability to accept proppant during fracturing
Generally, there are three to six perforation clusters for each frac- and a greater risk of proppant screenout. In summary, the key fea-
turing stage to create simultaneous multiple-fracture propagation tures for the multiple-fracture geometry are that the interior

–850 –850
FPM FPM
Wu et al. 2012 Wu et al. 2012

–900 –900

–950 –950

–1000 –1000
y, ft
y, ft

–1050 –1050

–1100 –1100

–1150 –1150

–1200 –1200
–60 –10 40 90 –60 –10 40 90
x, ft x, ft

Fig. 9—Comparison of propagation paths for two initially parallel fractures in isotropic and anisotropic stress fields: left figure
refers to isotropic stress field; right figure refers to anisotropic stress field.

342 April 2015 SPE Journal

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 343 Total Pages: 10

–850
–850 FPM
FPM
Wu et al. 2012
Wu et al. 2012

–900 –900

–950 –950

–1000 –1000

y, ft
y, ft

–1050 –1050

–1100 –1100

–1150 –1150

–1200 –1200
–60 –10 40 90 –60 –10 40 90
x, ft x, ft

Fig. 10—Comparison of propagation paths for two initially offset fractures in isotropic and anisotropic stress fields: left figure
refers to isotropic stress field; right figure refers to anisotropic stress field.

fractures are shorter, narrower, and more nonplanar than the exte- rxx is much larger than that for ryy, because fractures open against
rior fractures. the rock and directly squeeze the rock in the x-direction, whereas
One can further understand the resultant fracture paths and y-direction effects are reduced by Poisson’s ratio effects. Thus, at
widths by examining the stress distributions induced by the frac- some specific points the change in rxx is greater than the change
tures. Before fracturing, the minimum horizontal stress of rxx is ori- in ryy, such that rxx might become larger than ryy. This flip in the
ented parallel to the x-direction, and the maximum horizontal stress orientation of the compressive stress explains why interior frac-
is the ryy. Fractures grow along the y-direction because that path pro- tures change their propagation directions to seek the path with
vides the least resistance. As the fractures grow, geomechanical least resistance.
stresses around fractures are dynamically changed. In the Figs. 14 One also can use the induced stress field to predict the poten-
and 15, red colors indicate an increase of compressional stress, and tial for microseismic activity. Microseismic events are induced by
blue colors represent a decrease of compression. Nearby and ahead changes in stresses and pore pressure caused by hydraulic fractur-
of the fracture tips, compressional stresses of rxx and ryy decrease ing (Warpinski et al. 2001; Vulgamore et al. 2007), and are often
because of induced tensile stresses in the near-tip field. In contrast, used to estimate fracture geometry. The stress changes generated
rock behind the tips is compressed because of the fracture opening. by opening fractures might lead to the shear failure of pre-existing
In comparing the change in rxx and ryy at the end of injection fractures. Fig. 16 shows the change of the maximum shear stress
(Figs. 14 and 15), one can note that the area of stress change for at the end of fracturing. As shown in the figure, areas representing
0 psi of the shear stress change indicate no change from the initial
state. Red colors mean an increase in the maximum shear stress,
TABLE 4—INPUT PARAMETERS FOR CASE STUDY whereas blue colors indicate a decrease. The initial maximum
shear stress is 50 psi because the differential stress (SHmax–Shmin)
Fracture Spacing 50 ft is 100 psi. A larger maximum shear stress suggests the greater
Layer Height 100 ft likelihood to induce shear failure. Fig. 16 illustrates that the large
Minimum Horizontal Stress 4,450 psi increase of the shear-stress primarily occurs around fracture tips
Maximum Horizontal Stress 4,550 psi and the region between two fractures. Accordingly, in these areas,
Fluid Leakoff coefficient 0.00001 ft/min0.5
there is a possibility to induce shear failure of pre-existing fac-
tures and generate microseismic activity. Warpinski et al. (2001)
Injection Rate 60 bbl/min
reported that considerable shear stress near fracture tips developed
n0 0.7 by hydraulic fractures provides a primary mechanism for microse-
k0 0.02 lbf-secn/ft2 isms to accurately determine the length and height of fractures,
Young’s Modulus 6.53106 psi and pore-pressure change because of fluid leakoff is an important
Poisson’s Ratio 0.2 factor in microseismic development behind the tip. Our analysis,
Number of Perforations 60 however, did not take into account fluid-leakoff effects.
Density of Slurry 1.2 g/cm3
Density of Water 1.0 g/cm3
Conclusions
Diameter of Perforations 5 mm
The FPM presented in this paper is able to simulate simultaneous
Diameter of Wellbore 0.1 m
multiple-fracture propagation. The model integrates fracture

April 2015 SPE Journal 343

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 344 Total Pages: 10

Frac. 1 Frac. 2 Frac. 3

Frac. 4 Frac. 5 Frac. 6


500

400

300

200

100
y, ft

–100

–200

–300

–400

–500
–200 –100 0 100

x, ft
(a) Fracture trajectory (b) Fracture width distribution

Fig. 11—Fracture trajectory and aperture distribution after fracturing (map view), twofold exaggeration of x-axis scale.

interaction between multiple fractures and fluid flow in the single • Shear stress is generated during fracturing and might cause shear
fracture and the horizontal wellbore. Mass balance and pressure slippage of pre-existing fractures and develop microseismic ac-
equilibrium between multiple fractures and the horizontal well- tivity to help determine the geometry of multiple fractures.
bore are satisfied. From our model, we can conclude: • A map of the maximum shear stress change can assess the
• Simultaneous propagating fractures extend, attracting or repel- expected area of the shear failure of pre-existing fractures
ling each other as a result of stress-shadow effects, resulting in induced by hydraulic fracturing.
the direction and magnitude change of maximum and minimum
horizontal stresses.
• Under the same situation, fractures closer to the injection point Nomenclature
can obtain more fluid and propagate longer and wider. a ¼ half-length of the element, ft
• Nonplanar-geometry reduced propagation occurs for interior CL ¼ fluid-loss coefficient, ft/min0.5
fractures, causing a significant restriction of fracture aperture D ¼ diameter of wellbore, in.
and reduction of length. Ds ¼ shear displacement of the element, in.
• Compression beside the fractures increases, but is decreased at Dn ¼ normal displacement of the element, in.
fracture tips. At some points around fractures, the increment of dij ¼ distance between the center of elements i and j, ft
minimum horizontal stress is much larger than that of maxi- dp,i ¼ diameter of perforation, in.
mum horizontal stress, resulting in direction change of maxi- E ¼ Young’s modulus, psi
mum horizontal stress and nonplanar fracture geometry. E0 ¼ plane-strain modulus, psi

800
0.25 Frac. 1 Frac. 2 Frac. 3
Frac. 1 Frac. 2 700 Frac. 4 Frac. 5 Frac. 6
Frac. 3 Frac. 4
0.02 Frac. 5 Frac. 6 600
Net pressure, psi
Fracture width, in.

500
0.15
400
300
0.01
200
0.05 100
0
0 0 100 200 300 400 500
0 100 200 300 400 500
Fracture length, ft
Fracture length, ft
Fig. 13—Net pressure in each fracture changing along fracture
Fig. 12—Fracture width vs. fracture length. length.

344 April 2015 SPE Journal

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 345 Total Pages: 10

Change of normal stress in x-direction/psi Change of normal stress in y-direction/psi

500 300 500 300


400 400
200 200
300 300
200 200
100 100
100 100
y, ft

y, ft
0 0 0 0
–100 –100
–100 –100
–200 –200
–300 –200 –300 –200
–400 –400
–500 –300 –500 –300

–200 100 0 200 200 –200 100 0 200 200


x, ft x, ft

Fig. 14—Stress change of normal stress rxx (map view); twofold Fig. 15—Stress change of normal stress ryy (map view); twofold
exaggeration of x-axis scale. exaggeration of x-axis scale.

h ¼ fracture height, ft t ¼ injection time, minutes


i ¼ injection rate, bbl/min u
i ¼ displacement over the boundary S of the region, in.
KI ¼ stress-intensity factor of opening mode, psim0.5 uþ
i ¼ displacement over the boundary Sþ of the region, in.
KII ¼ stress intensity factor of shearing mode, psim0.5  ¼ Poisson’s ratio
k0 ¼ fluid-consistency index, lbf-secn/ft2 w ¼ fracture width, in.
Li(t) ¼ total fracture length of ith fracture at time t, ft a ¼ empirically determined constants, 2.0
N ¼ the number of fractures ar ¼ retardation coefficient
n0 ¼ fluid power-law index b ¼ empirically determined constants, 2.0
np,i ¼ the number of perforations DrI ¼ the mode I driving stress
p ¼ internal pressure, psi l ¼ fluid viscosity, cp
Pfrac ¼ pressure in the fractures, psi qs ¼ density of slurry, g/cm3
pk ¼ net pressure at the kth iteration step, psi qw ¼ density of water, g/cm3
q ¼ fluid-flow rate in the fracture, bbl/min sðsÞ ¼ time at which fracture reaches position s, minutes
Qi ¼ injection volume of each fracture, bbl/min h ¼ counterclockwise positive angle from the direction of
qL(s,t) ¼ volume rate of fluid loss to the formation per unit length in-plane propagation
of fracture, bbl/min
QT ¼ total injection rate, bbl/min References
s¼ distance along the fracture, ft
Shmax ¼ maximum horizontal stress, psi Adachi, J., Siebrits, E., Peirce, A. et al. 2007. Computer Simulation of Hy-
Shmin ¼ minimum horizontal stress, psi draulic Fractures. International J. Rock Mechanics & Mining Sci. 44:
739–757.
Carter, R.D. 1957. Appendix to Optimum Fluid Characteristics for Frac-
ture Extension. In Drill. and Prod. Prac., ed. G.C. Howard and C.R.
Change of maximum shear stress/psi Fast. API 267.
500 Castonguay, Stephen T., Mear, Mark E., Dean, Rick H. et al. 2013. Predic-
tions of the Growth of Multiple Interaction Hydraulic Fractures in
400 Three Dimensions. Presented at the SPE Annual Technical Conference
100
and Exhibition, New Orleans, Louisiana, 30 September–2 October.
300
SPE-166259-MS. http://dx.doi.org/10.2118/166259-MS.
200 50 Chong, K.K., Grieser, W.V., Passman, A. et al. 2010. A Completions
Guide Book to Shale-Play Development: A Review of Successful
100
Approaches Towards Shale-Play Stimulation in the Last Two Decades.
y, ft

0 0 Presented at the Canadian Unconventional Resources and International


Petroleum Conference, Calgary, Alberta, Canada, 19–21 October.
–100 CSUG/SPE-133874-MS. http://dx.doi.org/10.2118/133874-MS.
–200 –50 Crouch, S.L. 1976. Solution of Plane Elasticity Problems by the Displace-
ment Discontinuity Method. International J. Numerical Methods in
–300 Eng. 10: 301–343.
–400 –100 Crouch, S.L. and Starfield, A.M. 1983. Boundary Element Methods in
Solid Mechanics. London: George Allen & Unwin.
–500 Dahi_Taleghani, A. and Olson, J.E. 2011. Numerical Modeling of Multi-
–200 100 0 200 200 stranded-Hydraulic-Fracture Propagation: Accounting for the Interac-
x, ft tion Between Induced and Natural Fractures. SPE J. 16 (3): 575–581.
SPE-124884-PA. http://dx.doi.org/10.2118/124884-PA.
Fig. 16—The change of maximum shear stress (map view); two- Daneshy, A. 2011. Multistage Fracturing Using Plug-and-Perf Systems.
fold exaggeration of x-axis scale. Shale Energy/Fracturing 232 (10).

April 2015 SPE Journal 345

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J167626 DOI: 10.2118/167626-PA Date: 10-April-15 Stage: Page: 346 Total Pages: 10

Elbel, J.L., Piggott, A.R., and Mack, M.G. 1992. Numerical Modeling of Schlichting, H. 1968. Boundary-Layer Theory, sixth edition. New York:
Multilayer Fracture Treatments. Presented at the SPE Permian Basin McGraw-Hill.
Oil and Gas Recovery Conference, Midland, Texas, 18–20 March. Sheibani, Farrokh and Olson, Jon. 2013. Stress Intensity Factor Determi-
SPE-23982-MS. http://dx.doi.org/10.2118/23982-MS. nation for Three-Dimensional Crack Using the Displacement Disconti-
Erdogan, F. and Sih, G.C. 1963. On the Crack Extension in Plates Under nuity Method With Applications to Hydraulic Fracture Height Growth
Plane Loading and Transverse Shear. J. Basic Eng. 519–527. and Non-Planar Propagation Paths. In Effective and Sustainable Hy-
Gale, Julia F.W., Reed, Robert M., and Holder, Jon. 2007. Natural Frac- draulic Fracturing, ed. A.P. Bunger, J. McLennan, and R. Jeffrey,
tures in the Barnett Shale and Their Importance for Hydraulic Fracture Chap. 37. Rijeka, Croatia: InTech. http://dx.doi.org/10.5772/56308.
Treatments. AAPG Bull. 91 (4): 603–622. http://dx.doi.org/10.1306/ Shou, K.J. 1993. A High Order Three-Dimensional Displacement Discon-
11010606061. tinuity Method With Application to Bonded Half-Space Problems. PhD
Keshavarzi, R. and Mohammadi, S. 2012. A New Approach for Nu- thesis, University of Minnesota, Minnesota (August 1993).
merical Modeling of Hydraulic Fracture Propagation in Naturally Siriwardane, H.J. and Layne, A.W. 1991. Improved Model for Predicting
Fractured Reservoirs. Presented at the SPE/EAGE European Un- Multiple Hydraulic Fracture Propagation From a Horizontal Well. Pre-
conventional Resources Conference and Exhibition, Vienna, Aus- sented at the SPE Eastern Regional Meeting, Lexington, Kentucky,
tria, 20–22 March. SPE-152509-MS. http://dx.doi.org/10.2118/ 22–25 October. SPE-23448-MS. http://dx.doi.org/10.2118/23448-MS.
152509-MS. Spence, D.A. and Sharp, P.W. 1985. Self-Similar Solution for Elastohy-
Li, Yi. 1991. On Initiation and Propagation of Fractures From Deviated drodynamic Cavity Flow. Proc. of the Royal Society of London, Series
Wellbores. Dissertation, The University of Texas at Austin (May A (400), pp. 289–313.
1991). Valko, P. and Economides, M.J. 1995. Hydraulic Fracture Mechanics.
Mack, Mark G., Elbel, Jacques L., and Piggott, Andrew R. 1992. Numeri- New York: John Wiley & Sons.
cal Representation of Multilayer Fracturing. Rock Mechanics, ed. Till- Vulgamore, T., Clawson, T., Pope, C. et al. 2007. Applying Hydraulic
erson and Wawersik. Balkema, Rotterdam. ISBN 90 5410 0451. Fracture Diagnostics to Optimize Stimulations in the Woodford Shale.
Meyer, B.R. and Bazan, L.W. 2011. A Discrete Fracture Network Model Presented at the SPE Annual Technical Conference and Exhibition,
for Hydraulically Induced Fractures: Theory, Parametric and Case Anaheim, California, 11–14 November. SPE-110029-MS. http://
Studies. Presented at the SPE Hydraulic Fracturing Technology Con- dx.doi.org/10.2118/110029-MS.
ference and Exhibition, The Woodlands, Texas, 24–26 January. SPE- Warpinski, N.R. and Branagan, P.T. 1989. Altered-Stress Fracturing. J Pet
140514-MS. http://dx.doi.org/10.2118/140514-MS. Technol 41 (9): 990–997. SPE-17533-PA. http://dx.doi.org/10.2118/
Nagel, N., Gil, I., Sanchez-Nagel, M. et al. 2011. Simulating Hydraulic 17533-PA.
Fracturing in Real Fractured Rock—Overcoming the Limits of Pseudo Warpinski, N.R., Moschovidis, Z.A., Parker, C.D. et al. 1994. Comparison
3D Models. Presented at the SPE Hydraulic Fracturing Technology Study of Hydraulic Fracturing Models—Test Case: GRI Staged Field
Conference and Exhibition, The Woodlands, Texas, 24–26 January. Experiment No. 3. SPE Prod & Fac 9 (1): 7–16. SPE-28158-PA.
SPE-140480-MS. http://dx.doi.org/10.2118/140480-MS. Warpinski, N.R., Wolhart, S.L., and Wright, C.A. 2001. Analysis and Pre-
Olson, J.E. 1990. Fracture Mechanics Analysis of Joints and Veins. Dis- diction of Microseismicity Induced by Hydraulic Fracturing. Presented
sertation, Stanford University, Stanford, California (December 1990). at the SPE Annual Technical Conference and Exhibition, New Orle-
Olson, J.E. 2004. Predicting Fracture Swarms—The Influence of Subcriti- ans, Louisiana, 30 September–3 October. SPE J. 9 (1): 24–33. SPE-
cal Crack Growth and the Crack-Tip Process Zone on Joint Spacing in 87673-PA. http://dx.doi.org/10.2118/87673-PA.
Rock. In The Initiation, Propagation, and Arrest of Joints and Other Weng, X., Kresse, O., Cohen, C. et al. 2011. Modeling of Hydraulic Frac-
Fractures, ed. J.W. Cosgrove and T. Engelder, 73–87. Geological So- ture Network Propagation in a Naturally Fractured Formation. Pre-
ciety of London Special Publication 231. sented at the SPE Hydraulic Fracturing Technical Conference and
Olson, J.E. 2007. Fracture Aperture, Length and Pattern Geometry Devel- Exhibition, The Woodlands, Texas, 24–26 January. SPE-140253-MS.
opment Under Biaxial Loading: A Numerical Study With Applications http://dx.doi.org/10.2118/140253-MS.
to Natural, Cross-Jointed Systems. In Fracture-Like Damage and Wong, Sau-Wai and Xu, Guanshui. 2013. Interaction of Multiple Hydraulic
Localisation, ed. G. Couples and H. Lewis, Vol. 289, 123-142. Geo- Fractures in Horizontal Wells. Presented at the SPE Middle East Uncon-
logical Society of London, Special Publications. ventional Gas Conference and Exhibition, Muscat, Oman, 28–30 Janu-
Olson, J.E. 2008. Multi-Fracture Propagation Modeling: Applications to ary. SPE-163982-MS. http://dx.doi.org/10.2118/163982-MS.
Hydraulic Fracturing in Shales and Tight Gas Sands. Presented at the Wu, R., Kresse, O., Weng, X. et al. 2012. Modeling of Interaction of Hy-
42nd U.S. Symposium on Rock Mechanics, San Francisco, California, draulic Fractures in Complex Fracture Networks. Presented at the SPE
29 June–2 July. Hydraulic Fracture Technology Conference, The Woodlands, Texas,
Olson, J.E. and Wu, K. 2012. Sequential Versus Simultaneous Multi-zone 6–8 February. SPE-152052-MS. http://dx.doi.org/10.2118/152052-MS.
Fracturing in Horizontal Wells: Insights From a Non-planar, Multi- Zimmerman, R.W., Kumar, S., and Bodvarsson, G.S. 1991. Lubrication
frac Numerical Model. Presented at the SPE Hydraulic Fracturing Theory Analysis of the Permeability of Rough-Walled Fractures. Inter-
Technology Conference, The Woodlands, Texas, 6–8 February. SPE- national J. Rock Mechanics Mine Sci. & Geomechanics. Abstr. 28 (4):
152602-MS. http://dx.doi.org/10.2118/152602-MS. 325–331.
Olson, J.E. and Dahi Taleghani, A. 2009. Modeling Simultaneous Growth
of Multiple Hydraulic Fractures and Their Interaction With Natural Kan Wu is a PhD candidate in the Department of Petroleum and
Fractures. Presented at the SPE Hydraulic Fracturing Technology Con- Geosystems Engineering at the University of Texas at Austin (UT).
ference, The Woodlands, Texas, 19–21 January. SPE-119739-MS. Her research is focused on developing a fracture-propagation
http://dx.doi.org/10.2118/119739-MS. model to simulate multiple fracture propagation in naturally
Olson, J.E. and Pollard, D.D. 1989. Inferring Paleostresses From Natural fractured reservoirs. Wu holds MS and BS degrees in petroleum
Fracture Patterns: A New Method. Geology 17 (4): 345–348. http:// engineering from the China University of Petroleum. She is a
member of SPE. Wu won the First Place in the PhD Division at the
dx.doi.org/10.1130/0091-7613.
2013 SPE Gulf Coast Region Paper Contest and Second Place in
Pollard, D.D. and Segall, P. 1987. Theoretical Displacements and Stresses the PhD Division at the 2013 SPE Local Paper Contest.
Near Fractures in Rock: With Applications to Faults, Joints, Veins,
Dikes and Solution Surfaces. In Fracture Mechanics of Rock, ed. B.K. Jon E. Olson is an associate professor in petroleum and geosys-
Atkinson, 277-350. London: Academic Press. tems engineering at UT. He is co-principal investigator of the
Fracture Research and Application Consortium at UT, and spe-
Roussel, Nicolas P. and Sharma, Mukul M. 2011. Strategies to Minimize
cializes in rock-mechanics applications to structural geology,
Frac Spacing and Stimulate Natural Fractures in Horizontal Comple- hydraulic fracturing, and reservoir characterization. Before join-
tions. Presented at the SPE Annual Technical Conference and Exhibi- ing UT in 1995, Olson spent 6 years as a research engineer at
tion. Denver, Colorado, 30 October–2 November. SPE-146104-MS. Mobil Research and Development Corporation. He holds a
http://dx.doi.org/10.2118/146104-MS. PhD degree in geomechanics from Stanford University.

346 April 2015 SPE Journal

ID: jaganm Time: 11:59 I Path: S:/3B2/J###/Vol00000/140037/APPFile/SA-J###140037


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 845 Total Pages: 13

A Generalized Framework Model for the


Simulation of Gas Production in
Unconventional Gas Reservoirs
Yu-Shu Wu, SPE, Colorado School of Mines; Jianfang Li, SPE, Research Institute of Petroleum Exploration and
Development; Didier-Yu Ding, SPE, IFP Energies Nouvelles; Cong Wang, SPE, Colorado School of Mines; and
Yuan Di, Peking University

Summary Darcy flow behavior, adsorption/desorption, strong interactions


Unconventional gas resources from tight-sand and shale gas reser- between fluid (gas and water) molecules, and solid materials
voirs have received great attention in the past decade around the within tiny pores, as well as micro- and macrofractures of shale
world because of their large reserves and technical advances in and tight formations. Currently, there is little in basic understand-
developing these resources. As a result of improved horizontal- ing of how these complicated flow behaviors impact gas flow and
drilling and hydraulic-fracturing technologies, progress is being the ultimate gas recovery in such reservoirs. In particular, only a
made toward commercial gas production from such reservoirs, as few effective reservoir simulators and few modeling studies cur-
demonstrated in the US. However, understandings and technolo- rently are available (e.g., Kelkar and Atiq 2010) in the industry
gies needed for the effective development of unconventional res- for assisting reservoir engineers to model and develop the uncon-
ervoirs are far behind the industry needs (e.g., gas-recovery rates ventional natural-gas resources.
from those unconventional resources remain very low). There are Shale formations are characterized by extremely low perme-
some efforts in the literature on how to model gas flow in shale ability from subnanodarcies to microdarcies, and it is different for
gas reservoirs by use of various approaches—from modified com- different shale types, even under similar porosity, stress, or pore
mercial simulators to simplified analytical solutions—leading to pressure. As summarized by Wang et al. (2009), the permeability
limited success. Compared with conventional reservoirs, gas flow of deep organic-lean mudrocks ranges from smaller than to tens
in ultralow-permeability unconventional reservoirs is subject to of nanodarcies, whereas permeability values in organic-rich gas
more nonlinear, coupled processes, including nonlinear adsorption/ shales range from subnanodarcies to tens of microdarcies. The
desorption, non-Darcy flow (at both high flow rate and low flow Klinkenberg effect (Klinkenberg 1941), or gas-slippage effect,
rate), strong rock/fluid interaction, and rock deformation within has been practically ignored in conventional gas reservoir studies,
nanopores or microfractures, coexisting with complex flow geome- except when analyzing pressure responses or flow near gas-pro-
try and multiscaled heterogeneity. Therefore, quantifying flow in duction wells at a very low pressure. This is because of a larger
unconventional gas reservoirs has been a significant challenge, and the pore size and relatively high pressure existing in those conven-
traditional representative-elementary-volume- (REV) based Darcy’s tional gas reservoirs. In shale gas reservoirs, however, the Klin-
law, for example, may not be generally applicable. kenberg or slippage effect is expected to be significant because of
In this paper, we discuss a generalized mathematical framework the nanosized pores of such rock, even under a high-pressure con-
model and numerical approach for unconventional-gas-reservoir dition. Wang et al. (2009) show that gas permeability in the Mar-
simulation. We present a unified framework model able to incorpo- cellus shale increases from 19.6 md at 1,000 psi to 54 md at 80 psi
rate known mechanisms and processes for two-phase gas flow and because of the strong slippage effect.
transport in shale gas or tight gas formations. The model and numer- Unconventional reservoir dynamics are characterized by the
ical scheme are based on generalized flow models with unstructured highly nonlinear behavior of multiphase flow in extremely low-
grids. We discuss the numerical implementation of the mathematical permeability rock, coupled by many coexisting physical processes
model and show results of our model-verification effort. Specifically, (e.g., non-Darcy flow). Because of complicated flow behavior, a
we discuss a multidomain, multicontinuum concept for handling strong interaction between fluid and rock, and multiscaled hetero-
multiscaled heterogeneity and fractures [i.e., the use of hybrid mod- geneity, the traditional Darcy’s-law/REV-based model may not
eling approaches to describe different types and scales of fractures be generally applicable for describing flow phenomena in uncon-
or heterogeneous pores—from the explicit modeling of hydraulic ventional gas reservoirs. Blasingame (2008) and Moridis et al.
fractures and the fracture network in stimulated reservoir volume (2010) provide very comprehensive reviews of flow mechanisms
(SRV) to distributed natural fractures, microfractures, and tight ma- in unconventional shale gas reservoirs. Both studies point out that
trix]. We demonstrate model application to quantify hydraulic frac- the nonlaminar/non-Darcy flow concept of high velocity may turn
tures and transient flow behavior in shale gas reservoirs. out to be important in shale gas production. The nonlaminar/non-
Darcy flow concept of high-velocity flow in shale gas reservoirs
Introduction may not be represented by Darcy’s law, and the Forchheimer
Even with the significant progress made in producing natural gas equation is probably sufficient for many applications.
from unconventional, low-permeability shale gas and tight gas Natural gas in shale gas formations is present both as a free-gas
reservoirs in the past decade, gas recovery remains very low (esti- phase and as gas adsorbed onto solids in pores. In these reservoirs,
mated at 10 to 30% of gas in place). Gas production or flow in gas or methane molecules are adsorbed mainly to the carbon-rich
such extremely low-permeability formations is complicated fur- components (i.e., kerogen) (Silin and Kneafsey 2011; Mengal and
ther by many coexisting processes, such as severe heterogeneity, Wattenbarger 2011; EIA 2011). The adsorbed gas represents a sig-
a large Klinkenberg effect (Klinkenberg 1941), nonlinear or non- nificant percentage of total gas reserves (20 to 80%) as well as a sig-
nificant factor in recovery rates, which cannot be ignored in any
model or modeling analysis. In shale gas formations, past studies
found that methane molecules are adsorbed mainly to the carbon-
Copyright V
C 2014 Society of Petroleum Engineers
rich components (i.e., kerogen), correlated with total organic con-
This paper (SPE 163609) was accepted for presentation at the SPE Reservoir Simulation tent (TOC) in shales, as a function of reservoir pressure.
Symposium, The Woodlands, Texas, USA, 18–20 February 2013, and revised for
publication. Original manuscript received for review 4 March 2013. Revised manuscript
In conventional oil or gas reservoirs, the effect of geome-
received for review 13 January 2014. Paper peer approved 5 February 2014. chanics on rock deformation or permeability is generally small

October 2014 SPE Journal 845

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 846 Total Pages: 13

and has been mostly ignored in practice. However, in unconven- gaseous and aqueous. For simplicity, the gas and water compo-
tional shale formations with nanosized pores or nanosized micro- nents are assumed to be present only in their associated phases
fractures, such geomechanics effects can be relatively large and and adsorbed gas is within the solid phase of rock. Each fluid
may have a significant impact on both fracture and matrix perme- phase flows in response to pressure and gravitational and capillary
ability, which has to be considered, in general. Wang et al. (2009) forces according to the multiphase extension of Darcy’s law or
show that permeability in the Marcellus shale is pressure-depend- several extended non-Darcy-flow laws, discussed next. In an iso-
ent and decreases with an increase in the confining pressure (or thermal system containing two mass components, subject to mul-
total stress). The effect of confining pressure on permeability is tiphase flow and adsorption, two mass-balance equations are
caused by a reduction of porosity. Bustin et al. (2008) report the needed to fully describe the system, as described in an arbitrary
effect of stress (confining pressure) in Barnett, Muskwa, Ohio, flow region of a porous or fractured domain for flow of phase b
and Woodford shales and show that the degree of permeability (b ¼ g for gas and b ¼ w for water),
reduction with confining pressure is significantly higher in shales
than in consolidated sandstone or carbonate. @
ð/Sb qb þ mb Þ ¼ r  ðqb vb Þ þ qb : . . . . . . . . . . . . ð1Þ
This paper presents a generalized mathematical model and nu- @t
merical approach for unconventional-gas-reservoir simulation.
where / is the effective porosity of the porous or fractured media;
We present a unified framework model that is able to incorporate
Sb is the saturation of fluid b; qb is the density of fluid b; vb is the
many known mechanisms and processes for two-phase gas flow in
volumetric velocity vector of fluid b, determined by Darcy’s law
shale gas or tight gas formations with a continuum modeling
or non-Darcy-flow models, discussed next; t is time; mb is the
approach. The numerical scheme is based on generalized flow
adsorption or desorption mass term for the gas component per
models with unstructured grids. We discuss the numerical imple-
unit volume of formation; and qb is the sink/source term of phase
mentation of Klinkenberg effects, non-Darcy flow, gas adsorption,
(component) b per unit volume of formation.
and geomechanics effects into the mathematical model. In the nu-
merical-modeling examples, we apply geomechanics-coupled per-
meability to fractures, and apply the Klinkenberg effect to matrix Incorporation of Gas Adsorption and Desorption. The amount
media. This is because the Klinkenberg effect is apparent with of adsorbed gas in a given shale gas formation is generally
extremely low permeability or tiny pores, whereas fractures are described with the Langmuir’s isotherm (e.g., Moridis et al. 2010;
very sensitive to the change in effective stress. Results of our Mengal and Wattenbarger 2011; Silin and Kneafsey 2011; EIA
model-verification effort are also presented. We demonstrate 2011; Wu et al. 2012; Wu and Wang 2012) (i.e., it is correlated to
model application to quantify hydraulic fractures and transient reservoir gas pressure). To incorporate the gas adsorption or de-
flow behavior in shale gas reservoirs. sorption mass term in the mass-conservation equation, the amount
One of the critical issues in shale-gas-reservoir simulation is of adsorbed gas is determined according to the Langmuir’s iso-
how to handle fracture flow and fracture/matrix interaction. This therm as a function of reservoir pressure. As the pressure
is because the gas flow and production rely on fractures in these decreases with continuous gas production through production
reservoirs. Cipolla et al. (2009) built a methodology on modeling wells in reservoirs, more adsorbed gas is released from the solid
complex fracture geometry and heterogeneity from the microseis- to the free-gas phase in the pressure-lowering region, contributing
mic data. In this paper, we present a hybrid fracture-modeling to the total gas flow or production. In our model, the mass of
approach, defined as a combination of explicit fracture modeling adsorbed gas in unit formation volume is described (Leahy-Dios
and multicontinuum, multiple-interacting-continua (MINC) (Pruess et al. 2011; Silin and Kneafsey 2011; Wu et al. 2012) as
and Narasimhan 1985), and single-porosity modeling approaches,
which seems the best option for modeling a shale gas reservoir mg ¼ qR qg VE ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ
with both hydraulic fractures and natural fractures. This is because
hydraulic fractures, which have to be dealt with for shale gas pro- where mg is adsorbed gas mass in unit formation volume; qR is
duction, are better handled by the explicit fracture method but can- rock bulk density; qg is gas density at standard condition; and VE
not be modeled, in general, by a dual-continuum model. On the is the adsorption isotherm function or gas content in scf/ton
other hand, naturally fractured reservoirs are better modeled by a (or standard gas volume adsorbed per unit rock mass). If the
dual-continuum approach, such MINC, for extremely low-perme- adsorbed-gas terms can be represented by the Langmuir isotherm
ability matrix in shale gas formations, which cannot be modeled by (Langmuir 1916), the dependency of adsorbed-gas volume on
an explicit fracture model. Specifically, we demonstrate how to use pressure at constant temperature is given as
the hybrid modeling approach to describe different types and scales
of fractures from the explicit modeling of hydraulic fractures and P
fracture network in the SRV to distributed natural fractures, micro- VE ¼ VL ; . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
P þ PL
fractures, and tight matrix.
where VL is the Langmuir’s volume in scf/ton; P is reservoir gas
pressure; and PL is Langmuir’s pressure, the pressure at which
Flow-Governing Equations 50% of the gas is desorbed. In general, Langmuir’s volume VL is
In most cases of gas production from shale gas formations, a two- a function of the organic richness (or TOC) and thermal maturity
phase (gas/liquid) -flow model or a multiphase-flow model is con- of the shale.
sidered to be sufficient for simulation studies. This is because Note that Eq. 3 is valid only for the case when the Langmuir
what we are most concerned with in shale-gas-reservoir simula- model is applicable. In general, VE in Eq. 2 can be determined
tion is the modeling of gas flow from reservoir to well. However, from any correlation of gas adsorption as a function of reservoir
in addition to the gas phase, liquid-phase flow is often occurring gas pressure, which may be defined by a table lookup from labora-
simultaneously with gas flow; it needs to be considered when two tory studies for a given unconventional reservoir.
cases exist—mobile in-situ connate water and an abundance of In the literature, the most commonly used empirical model
aqueous hydraulic-fracturing fluids, which are sucked into the for- describing sorption onto organic carbon in shales is analogous to
mations surrounding the wells. Therefore, in this paper, we pri- that used in coalbed methane and follows the Langmuir isotherm
marily discuss the two-phase (gas and liquid) -flow model and (Gao et al. 1994; Moridis et al. 2010), such as Eq. 2. This adsorp-
formulation and treat single-phase gas flow as a special case of tion-modeling approach is based on the assumption that an instan-
the two-phase flow for the simulation studies of unconventional taneous equilibrium exists between the sorbed and the free gas
gas reservoirs. (i.e., there is no transient-time lag between pressure changes and
A multiphase system of gas and water (or liquid) in a porous the corresponding sorption/desorption responses; i.e., the equilib-
or fractured unconventional reservoir is assumed to be similar to rium model of the Langmuir sorption is assumed to be valid,
what is described in a black-oil model, composed of two phases: which provides a good approximation in shale gas modeling).

846 October 2014 SPE Journal

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 847 Total Pages: 13

1E–01 where /0 is zero-effective-stress porosity; /r is high-effective-


stress porosity; and the exponent a is a parameter. Rutqvist et al.
(2002), Davies and Davies (1999), and Winterfeld and Wu (2011,
Gas Permeability, md
1E–02 2012) also present an associated function for permeability in
terms of porosity,
1 microdarcy
 
/
1E–03 c 1
k ¼ ko e /0 ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð10Þ
where c is a parameter. Fig. 1 shows the effect of confining pres-
1E–04 Barnett Muskwa sure on gas permeability in gas shales.
Ohio Marcellus One can use an alternative, table-lookup approach for the cor-
Woodford-1 Woodford-2
Huron relation of reservoir porosity and permeability as a function of
1E–05 effective mean stress, from laboratory studies, for a given uncon-
0 1600 3200 4800 6400 ventional reservoir.
Effective Stress, psi One must note that Fig. 1 from Wang et al. (2009) presents the
permeability measurement from core plugs in which potential natu-
Fig. 1—Effect of confining pressure on gas permeability in gas ral microfractures in core plugs play an important role for the con-
shales (Wang et al. 2009). nectivity. If one uses crushed samples to measure the matrix
permeability only by eliminating natural and drilling induced
Several kinetic sorption models exist in the literature that use dif- microfactures, the permeability value is one or two orders lower.
fusion approaches; however, the subject has not been fully inves- The geomechanics has a much stronger impact on the fracture than
tigated or fully understood (Moridis et al. 2010). on the matrix. So, when using a dual-porosity approach in the mod-
eling, if microfractures are considered as a part of the matrix media,
one can directly apply the relations in Eqs. 4 through 10. However,
Coupled Flow and Geomechanics Effect. In this section, we
if microfractures are considered as a part of the fractured media,
will propose a simple-to-implement modeling approach, easy to
the geomechanics effect is more complex because fracture conduc-
incorporate into an existing reservoir simulator, to couple geome-
tivities are subjected to different laws according to microfractures,
chanics with two-phase flow in unconventional reservoirs. The
partially propped fractures, or propped fractures (Cipolla 2009).
following discussion is based on our previous work (e.g., Wu
The applicability of these mechanics-coupling models in mul-
et al. 2008; Winterfeld and Wu 2011). The effective porosity, per-
tiphase-flow simulations for a rock-deformation effect requires
meability, and capillary pressure of rock are assumed to correlate
that the initial distribution of effective stress or total stress field be
with the mean effective stress (r0m ), defined as
predetermined as a function of spatial coordinates and pressure
r0m ¼ rðx; y; z; PÞ  aP; . . . . . . . . . . . . . . . . . . . . . . ð4Þ fields, as in Eq. 5. In practice, the stress distribution may be esti-
mated analytically, numerically, or from field measurements
where a is the Biot constant and because changes in effective stress are primarily caused by changes
in reservoir pressure during production. These models can be sig-
rm ðx; y; z; PÞ ¼ ½rx ðx; y; z; PÞ þ ry ðx; y; z; PÞ þ rz ðx; y; z; PÞ=3; nificantly simplified for coupling multiphase gas flow with rock de-
formation in stress-sensitive formations in numerical simulation, if
                   ð5Þ the in-situ total stress in reservoirs is nearly constant or a function
where rx, ry, and rz are total stress in the x-, y-, and z-direction, of spatial coordinates as well as fluid pressure only during the
respectively. With the definition of the mean effective stress in production. The constant-total-stress requirement may be approxi-
Eq. 5, the effective porosity of the formation (fractures or porous mately satisfied for deep reservoirs.
media) is defined as a function of mean effective stress only,
Incorporation of Klinkenberg or Gas-Slippage Effect. In low-
/ ¼ /ðr0m Þ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ permeability shale gas formations with nanosized pores or under a
low-reservoir-pressure condition, the Klinkenberg effect (Klin-
Similarly, the intrinsic permeability is related to the effective kenberg 1941) may be significant and should be accounted for
stress; that is, when modeling gas flow in such reservoirs (Wu et al. 1998; Wang
et al. 2009). As discussed previously, the Klinkenberg effect is
k ¼ kðr0m Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ expected to be larger or stronger in unconventional reservoirs
because of small pore size and low permeability in comparison
For capillary pressure functions, the impact of rock deforma- with those in conventional reservoirs. The Klinkenberg effect, if
tion or pore change is accounted for with the Leverett function existing, will enhance gas permeability or productivity in a low-
(Leverett 1941), pressure zone, such as the region near a well or matrix portions
qffiffiffiffiffiffiffiffiffiffiffiffi near fractures, of low-permeability unconventional formations,
and, therefore, it should be included as an additional beneficial
k0 =/0
Pc ¼ Cp P0c ðSw Þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; . . . . . . . . . . . . . . . . ð8Þ factor of gas-flow enhancement.
kðr0m Þ=/ðr0m Þ The Klinkenberg effect is incorporated in gas-flow models by
modifying absolute permeability for the gas phase as a function of
where Pc is the capillary pressure between gas and water as a gas pressure (e.g., Wu et al. 1998),
function of water or gas saturation; Cp is a constant; and the  
superscript 0 denotes reference or zero-stress condition. b
kg ¼ k1 1 þ ; . . . . . . . . . . . . . . . . . . . . . . . . ð11Þ
Several correlations have been used for porosity as a function Pg
of effective stress and permeability as a function of porosity (Da-
where k1 is constant, absolute gas-phase permeability under very
vies and Davies 1999; Rutqvist et al. 2002; Winterfeld and Wu
large gas-phase pressure (in which the Klinkenberg effect is mini-
2011, 2012). In our numerical implementation, the function for
mized); and b is the Klinkenberg beta factor and could be pressure-
porosity and permeability presented by Rutqvist et al. (2002) is
or temperature-dependent, accounting for the gas-slippage effect.
adopted, which is obtained from laboratory experiments on sedi-
In a conventional-gas-reservoir simulation, the beta factor is
mentary rock (Davies and Davies 1999),
commonly treated as constant and depends on the pore structure
0
/ ¼ /r þ ð/0  /r Þear ; . . . . . . . . . . . . . . . . . . . . . ð9Þ of the medium and formation temperature for a particular reser-
voir. Several recent studies on dynamic gas slippage with

October 2014 SPE Journal 847

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 848 Total Pages: 13

60 60

Permeability, µD

Permeability, µD
40 40

20 20

0 0
0 1000 2000 3000 0 0.005 0.01 0.015
Pore Pressure, psia Reciprocal of Pore Pressure, 1/psia

Fig. 2—Contribution of the Klinkenberg effect to the apparent matrix permeability (Ozkan et al. 2010).

microscale or pore-scale models have considered the beta factor We analyze the Klinkenberg effect with three different matrix
as a function of gas pressure or the Knudsen number. In applica- permeabilities—1:0  103 md, 1:0  105 md; and 1:0  107
tion, the Klinkenberg effect should be modeled with the labora- md; as shown in Fig. 2 and Fig. 4. We can see that the contribu-
tory-determined beta factor either as a constant or as a pressure- tion of the Klinkenberg effect is more significant at low pressures
dependent function or simply treating the apparent gas permeabil- and for lower values of permeability. This estimation also pro-
ity as a function of pressure from a table lookup to include the vides reliable values of the beta factor for analyzing the Klinken-
Klinkenberg effect or the Knudsen diffusion. An example relation berg effect.
between permeability and pressure, as shown in Fig. 2, can be
directly used for the reservoirs concerned, if a site-specific study The Incorporation of NonDarcy Gas Flow. In addition to mul-
provides such correlations or plots. tiphase Darcy flow, non-Darcy flow may also occur between and
After a comparison of Fig. 1 and Fig. 3, the Klinkenberg effect among the continua, such as along fractures, in unconventional
seems to have less impact than that of geomechanics, and they are gas reservoirs. The flow velocity, vß, for the non-Darcy flow of
going in the opposite directions. Geomechanics has an effect each fluid may be described with the multiphase extension of the
mainly on the microfractures and stimulation fractures, whereas Forchheimer equation (e.g., Wu 2002),
the Klinkenberg effect is primarily on the matrix media with
nanosized pores or fractures. lb
ðrUb Þ ¼ vb þ bb qb vb jvb j; . . . . . . . . . . . . . ð14Þ
A dynamic gas-slippage theory was proposed on the basis of k  krb
the assumption that gas travels under the influence of a concentra-
tion field (random molecular flow) and a pressure field (macro- where ßß is presented the effective non-Darcy-flow coefficient
scopic flow) (Ertekin et al. 1986). According to this theory, the with a unit m–1 for fluid b under multiphase-flow conditions. The
Klinkenberg factor is not a constant anymore, but a pressure-de- correlation proposed by Evans and Civan is used to determine the
pendent value. They gave the expression of the Klinkenberg fac- non-Darcy-flow beta factor in the Forchheimer equation (Evans
tor as and Civan 1994) in our simulation examples, such as

Pcg Dg l 1:485  109


b¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð12Þ b¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . ð15Þ
kg k1:021

where cg is gas compressibility and D is diffusivity coefficient. where the unit of k is md and the unit of b is ft1 . This correlation
In Eq. 12, the correlation to compute the diffusivity constant is for b matched with more than 180 data points, including those for
given by Ertekin et al. (1986): propped fractures (correlation coefficient ¼ 0.974).

31:57 Numerical Model


Dg ¼ pffiffiffiffiffiffi k0:67 ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð13Þ
Mg As discussed previously, the partial-differential equation that gov-
erns gas and liquid flow in shale gas reservoirs is nonlinear. In
where Dg is in ft2 =D. addition, gas flow in unconventional reservoirs is subject to many

1.0E–2
3200
1.0E–3 km = 1E–7 md
km = 1E–3 md
1600
Permeability, md

1.0E–4
b factore, psi

1.0E–5 800
km = 1E–5 md km = 1E–5 md
1.0E–6 400

1.0E–7 200
km = 1E–7 md
1.0E–8 km = 1E–3 md
100
300.0 600.0 1200.0 2400.0 4800.0
300.0 600.0 1200.0 2400.0 4800.0
Pressure, psi
Pressure, psi
Fig. 3—Effect of pore pressure on gas permeability in the Mar-
cellus shale, with a confining pressure of 3,000 psi (Soeder Fig. 4—Estimations of Klinkenberg beta factor for three perme-
1988; Wang et al. 2009). ability values.

848 October 2014 SPE Journal

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 849 Total Pages: 13

other nonlinear flow processes, such as adsorption and non-Darcy In this numerical approach, we apply the upstream weighting
flow. In general, the flow model needs to be solved with a numeri- method to the mobility term and the harmonic mean method to
cal approach. This work follows the methodology for reservoir the transmissivity term to guarantee the convergence and accu-
simulation (i.e., the use of numerical approaches to simulate gas racy of the calculation. The flow-potential term in Eq. 17 is
and water flow), following three steps: (1) spatial discretization of defined as
mass-conservation equations; (2) time discretization; and (3) iter-
ative approaches to solve the resulting nonlinear, discrete alge- Ub j ¼ Pbi  qb;ijþ1=2 gZi ; . . . . . . . . . . . . . . . . . . . . . ð20Þ
braic equations.
where Zi is the depth to the center of block i from a reference
Discrete Equations. The component mass-balance equations datum.
(Eq. 1) are discretized in space with a control-volume or inte-
grated finite-difference concept (Pruess et at. 1999). The control- Handling the Klinkenberg Effect. To include the Klinkenberg
volume approach provides a general spatial discretization scheme effect on gas flow, the absolute permeability to gas phase in Eq.
that can represent a 1D, 2D, or 3D domain with a set of discrete 19 should be evaluated with Eq. 11 as a function of gas-phase
meshes. Each mesh has a certain control volume for a proper pressure.
averaging or interpolation of flow and transport properties or ther-
modynamic variables. Time discretization is carried out with a Handling the Non-Darcy Flow. Under the non-Darcy-flow con-
backward, first-order, fully implicit finite-difference scheme. The dition of Eq. 14, the flow term (flowbij ) in Eq. 17 along the con-
discrete nonlinear equations for components of gas and water at
nection (i, j), between elements i and j, is numerically defined as
gridblock or node i can be written in a general form:
(Wu 2002)
h i 8
b;n Vi 2 !2 31=2 9
ð/qSÞb;nþ1
i þ mb;nþ1
i  ð/qSÞb;n
i  mi >
< >
=
X Dt Aij 1 1
flowb;ij ¼  þ4  c ij ðUb j  Ub i Þ5 ;
¼ flowb;nþ1
ij þ Qb;nþ1
i 2ðkbb Þijþ1=2 >
: kb kb >
;
j 2 gi

ðb ¼ gas and liquidÞ and ði¼ 1; 2; 3;…;NÞ;                    ð21Þ


                   ð16Þ in which the non-Darcy-flow transmissivity is defined as
where superscript b serves also as an equation index for gas and 4ðk2 qb bb Þijþ1=2
water components, with b ¼ 1 (gas) and b ¼ 2 (water); superscript c ij ¼ : . . . . . . . . . . . . . . . . . . . . . . . ð22Þ
n denotes the previous time level, with n þ 1 as the current time Di þ D j
level to be solved; subscript i refers to the index of gridblock or
In evaluating the “flow” terms in Eqs.17 through 22, subscript
node i, with N as the total number of nodes in the grid; Dt is time- ij þ 1/2 is used to denote a proper averaging or weighting of
step size; Vi is the volume of node i; gi contains the set of direct
fluid-flow properties at the interface or along the connection
neighboring nodes ( j) of node i; and mki , flowkij , and Qki are the between two blocks or nodes i and j. For example, we use
adsorption or desorption, the component mass “flow” term upstream weighting for relative permeability, density, and non-
between nodes i and j, and sink/source term at node i for compo-
Darcy coefficient. The convention for the signs of flow terms is
nent k, respectively. that flow from node j into node i is defined as “þ” (positive) in
Eq. 16 presents a precise form of the balance equation for each calculating the flow terms.
mass component of gas and water in a discrete form. It states that
the rate of change in mass accumulation (plus adsorption or desorp-
tion, if existing) at a node over a timestep is exactly balanced by an Handling Fractured Media. Handling flow through fractured
inflow/outflow of mass and also by sink/source terms, when exist- media is critical in shale-gas-reservoir simulation, because gas
ing for the node. As long as all flow terms have the flow from node production from such low-permeability formations relies on frac-
i to node j equal to and opposite to that of node j to node i for fluids, tures, from hydraulic-fracture networks to various-scaled natural
no mass will be lost or created in the formulation during the solu- fractures, to provide flow channels for gas flow into producing
tion. Therefore, the discretization in Eq. 16 is conservative. wells. Therefore, any unconventional reservoir simulator must
The “flow” terms in Eq. 16 are mass fluxes by advective proc- have the capability of handling fractured media. The published
esses and are described, when Darcy’s law is applicable, by a dis- modeling exercises in the literature have paid much attention to
crete version of Darcy’s law; that is, the mass flux of fluid phase b modeling fractures in shale gas formations (e.g., Cipolla 2009;
along the connection is given by Freeman et al. 2009a,2009b; 2010; Moridis et al. 2010; Cipolla
et al. 2010; Rubin 2010; Li et al. 2011; Wu et al. 2012). However,
flowbij ¼ kb;ijþ1=2 cij ðUbj  Ubi Þ; . . . . . . . . . . . . . . . . ð17Þ note that very few studies have been carried out to address the
critical issues as how to accurately simulate fractured unconven-
where kb,i jþ1/2 is the mobility term to phase b, defined as tional gas reservoirs or to select the best approach for modeling a
given shale gas formation. Most of the modeling exercises use
! commercial reservoir simulators, developed for conventional-
qb krb
kb;ijþ1=2 ¼ : . . . . . . . . . . . . . . . . . . . . ð18Þ fractured-reservoir simulation, which have very limited capabil-
lb ities for modeling multiscaled or complicated fractured reservoirs.
ijþ1=2
On the other hand, to simulate fractured unconventional gas reser-
In Eq. 17, cij is transmissivity and is defined, for a Voronoi voirs, more effort on model development is needed—from new
grid, as (Pruess et al. 1999) conceptual models to in-depth modeling studies of laboratory to
field-scale application.
Aij kijþ1=2 In our opinion, the hybrid fracture-modeling approach—
cij ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . ð19Þ defined as a combination of explicit fracture modeling (discrete-
Di þ D j
fracture model) and MINC (Pruess and Narasimham 1985; Pruess
where Aij is the common interface area between the connected 1983) and single-porosity modeling approaches—provides the
blocks or nodes i and j; Di is the distance from the center of block best option for modeling a shale gas reservoir with both hydraulic
i to the common interface of blocks i and j; and kijþ1/2 is an aver- fractures and natural fractures. This is because hydraulic fractures,
aged (such as harmonic-weighted) absolute permeability along which have to be dealt with for shale gas production, are better
the connection between elements i and j. handled by the explicit fracture method, and they cannot be

October 2014 SPE Journal 849

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 850 Total Pages: 13

Fig. 5. Therefore, the MINC method treats interporosity flow in a


fully transient manner by computing the gradients that drive inter-
Fractures porosity flow at the matrix/fracture interface. In comparison with
the double-porosity or dual-permeability model, MINC does not
rely on the pseudosteady-state assumption to calculate fracture/
matrix flow and is able to simulate fully transient fracture/matrix
Matrix Blocks
interaction by subdividing nested-cell gridding inside matrix
blocks. The MINC concept should be generally applicable for
handling fracture/matrix flow in fractured-shale gas reservoirs, no
matter how large the matrix-block size is or how low the matrix
permeability is, and it is more suitable for handling fractured-
shale gas reservoirs. However, the MINC approach may not be
applicable to systems in which fracturing is so sparse that the frac-
tures cannot be approximated as a continuum.
As Fig. 6 shows, in our hybrid fracture model, both the hy-
draulic fractures and SRV are evaluated from the microseismic
cloud. Recent advances in microseismic-fracture mapping tech-
nology have provided previously unavailable information to char-
acterize hydraulic-fracture growth and SRV, and have documented
Fig. 5—Schematic of MINC concept (Pruess and Narasimham surprising complexities in many geological environments. We will
1985). have a primary hydraulic-fracture system and an associated stimu-
lated volume in each hydraulic-fracture stage. First, we define a pri-
mary fracture on the basis of the orientation and region of the
modeled, in general, by a dual-continuum model. On the other microseismic cloud. The hydraulic fractures are modeled by the
hand, naturally fractured reservoirs are better modeled by a dual- discrete-fracture method. We assume the SRV near the hydraulic
continuum approach, such as MINC, for extremely low-perme- fractures is the region with natural fractures, and we apply MINC
ability matrix in shale gas formations, which cannot be modeled in this region. Single-porosity is applied in the region outside the
by an explicit-fracture model. SRV, in which there are no natural fractures. Local grid refinement
An explicit-fracture-modeling, or discrete-fracture, concept is (LGR) is used to improve simulation accuracy because pressure
to include every fracture explicitly in the modeled system by the gradients change substantially over short distances in the regions
use of refined grids to discretize fractures and the matrix surround- near hydraulic fractures. LGR is performed near the hydraulic-frac-
ing fractures. This approach is a good option for simulating hy- ture region.
draulic fractures for gas production from hydraulic-fractured wells
in a nonfractured/shale gas reservoir. The advantage of this
approach is that it can model hydraulic fractures accurately when Numerical Solution. In this work, we use the fully implicit
the fractures are known for their spatial distributions, determined scheme to solve the discrete nonlinear Eq. 16 with a Newton itera-
from other fracture-characterization studies. The disadvantage is tion method. Let us write the discrete nonlinear equation, Eq. 16,
that it cannot be used for simulating natural fractures or microfrac- in a residual form as
tures, in general, because the number of natural fractures or micro- h i
b;n Vi
fractures in a shale gas reservoir is too large for the model to Rb;nþ1
i ¼ ð/qSÞb;nþ1
i þ mb;nþ1
i  ð/qSÞb;n
i  mi
handle and their actual distributions in formations are unknown. X Dt
For the low matrix permeability or large matrix-block size, the  flowb;nþ1
ij  Q b;nþ1
i ¼ 0
traditional double-porosity model may not be applicable for mod- j 2 gi
eling natural fractures in unconventional reservoirs. This is ðb ¼ 1; 2; i¼ 1; 2; 3;…; NÞ:           ð23Þ
because it takes years to reach the pseudosteady state under which
the double-porosity model applies. The MINC concept (Pruess Eq. 23 defines a set of 2  N coupled nonlinear equations that
and Narasimham 1985) is able to describe gradients of pressures, need to be solved for every balance equation of mass components,
temperatures, or concentrations near the matrix surface and inside respectively. In general, two primary variables per node are
the matrix—by further subdividing individual matrix blocks with needed to use the Newton iteration for the associated two equa-
1D or multidimensional strings of nested meshes, as shown in tions per node. The primary variables selected are gas pressure

Hydraulic fracture geometry LGR mesh

Microseismic
cloud
Dual-porosity, or dual-permeability, or
Well
MINC

Hydraulic fracture and SRV Discrete fracture Single porosity

Slide view of this model

Fig. 6—Hybrid fracture model built methodology from microseismic cloud.

850 October 2014 SPE Journal

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 851 Total Pages: 13

1.2E+7 8.E+05
Analytical c_β = 0
Analytical b = 7.6E6 Analytical c_β = 1.0E–3
Numerical b = 7.6E6 7.E+05 Analytical c_β = 1.0E–4
1.0E+7 Numerical b = 0 Numerical c_β = 0
Numerical c_β =1.0E–3
Analytical b = 0 6.E+05 Numerical c_β =1.0E–4

Pressure, Pa
8.0E+6

Pressure, Pa
5.E+05
6.0E+6 4.E+05

4.0E+6 3.E+05
2.E+05
2.0E+6
1.E+05
0.0E+0 0.E+00
0 2 4 6 8 10 0 2 4 6 8 10
Horizontal Distance, m Horizontal Distance, m

Fig. 7—Analytical and numerical results for linear flow with the Fig. 8—Analytical and numerical results for linear non-Darcy
Klinkenberg effect. flow.

and gas saturation. The rest of the dependent variables—such as steady and transient analytical solutions are derived or used for
relative permeability, capillary pressures, viscosity and densities, considering these flow mechanisms. The problem concerns steady-
adsorption term, and nonselected pressure and saturation—are state and transient gas flow across a 1D reservoir. The system con-
treated as secondary variables, which are calculated from selected tains steady-/transient-state gas flow at an isothermal condition, and
primary variables. a constant gas mass injection/production rate is imposed at one side
In terms of the primary variables, the residual equation, Eq. of the rock or well. The other boundary of the rock/reservoir is kept
23, at a node i is regarded as a function of the primary variables at at constant pressure. Eventually, the system will reach steady state,
not only node i, but also at all its direct neighboring nodes j. The if the production is maintained for a long period of time. A compar-
Newton iteration scheme gives rise to ison of the pressure profiles along the rock block from the simula-
tion and the analytical solution is shown in Figs. 7 and 8,
X @Rb;nþ1 ðxm;p Þ indicating that our simulated pressure distribution is in excellent
i
ðdxm;pþ1 Þ ¼ Rb;nþ1
i ðxm;p Þ; . . . . . . ð24Þ agreement with the analytical solutions for all the problems of 1D
m
@xm
linear flow with the Klinkenberg or non-Darcy-flow effect.
where xm is the primary variable m with m ¼ 1 and 2, respectively, Details about the analytical solution derivation considering the
at node i and all its direct neighbors; p is the iteration level; and Klinkenberg and non-Darcy-flow effect are included in our previ-
i ¼ 1, 2, 3, …, N. The primary variables in Eq. 23 need to be ous work (Wu et al. 2012), and we will show their verification
updated after each iteration, results only in this section for the 1D linear-flow steady-flow situa-
tion. Comparisons between the analytical and numerical solutions
xm;pþ1 ¼ xm;p þ dxm;pþ1 : . . . . . . . . . . . . . . . . . . . . . ð25Þ for the radial-flow and transient-flow cases are also presented in
our former work. Constant coefficients for the Klinkenberg effect
The Newton iteration process continues until the residuals and correlation (Eq. 13) for the non-Darcy-flow coefficient are used
Rb;nþ1
i or changes in the primary variables dxm;pþ1 over iteration with comparison results shown in Figs. 7 and 8.
are reduced below preset convergence tolerances.
Numerical methods are generally used to construct the Jaco- Verification for Flow With Adsorption. For the gas flow with
bian matrix for Eq. 24, as outlined in Forsyth et al. (1995). At adsorption, the approximate analytical solution is given in Appen-
each Newton iteration, Eq. 24 represents a system of (2  N) line- dix A. The parameters used for this comparison study are porosity
arized algebraic equations with sparse matrices, which are solved U ¼ 0.15; permeability k ¼ 100 md; formation temperature
by a linear equation solver. T ¼ 25 C; gas viscosity l ¼ 1:64  102 cp; initial pressure
Pi ¼ 105 Pa; and thickness of the radial system is 1 m. The
Numerical-Model Verification well-boundary condition is a constant gas/mass-injection rate:
To examine the accuracy of our simulator formulation in simulat- Q ¼ 1:0  104 kg=s.
ing porous-medium gas flow with the Klinkenberg, non-Darcy- Fig. 9 presents the comparisons of the pressure profile at 1.67
flow, gas-adsorption, and geomechanics effects, several relevant days from the numerical and analytical solutions. Two situations,

200000 200000
P_analytical
P_analytical
180000 180000 P_numerical
P_numerical
Pressure, Pa
Pressure, Pa

160000 160000

140000 140000

120000 120000
VL = 0 VL = 50
t = 1.67 days t = 1.67 days
100000 100000
0 0.5 1 1.5 2 2.5 0 1 2 3
Radius, m Radius, m

Fig. 9—Comparison of gas-pressure profiles considering gas adsorption in a radial system at 1.67 days, calculated with the numer-
ical and analytical solutions.

October 2014 SPE Journal 851

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 852 Total Pages: 13

3.0E+7
TABLE 1—PARAMETERS FOR CHECKING INTEGRAL
SOLUTION FOR FLOW WITH GEOMECHANICS EFFECT 2.5E+7

Parameter Value Unit

Pressure, Pa
2.0E+7
7
Initial pressure Pi ¼ 10 Pa
Initial porosity Ui ¼ 0:20 1.5E+7
3 Constant K Pressure (N)
Initial fluid density qw ¼ 975:9 kg=m
1.0E+7 Constant K Pressure (A)
Cross-sectional area a ¼ 1.0 m2 unconstant K Pressure (A)
Formation thickness h ¼ 1.0 m 5.0E+6 unconstant K Pressure (N)
Fluid viscosity l ¼ 0:35132  103 Pa  s
Fluid compressibility Cf ¼ 4:556  1010 Pa1 0.0E+0
Rock compressibility Cr ¼ 5:0  109 Pa1 0 5000 10000
Initial permeability k0 ¼ 9:860  1013 m2 Time, seconds
Water-injection rate qm ¼ 0:01 kg=s
Fig. 10—Comparison of injection pressures calculated from in-
Hydraulic radius rw ¼ 0:1 m
tegral and numerical solutions for linear flow in a permeability-
Exponential index c ¼ 2.22 dependent medium with constant and nonconstant permeabil-
ity function.

Langmuir volume VL ¼ 0 and VL ¼ 50m3 =kg, are considered.


The analytical solutions give an excellent match with the numeri- tures, and therefore opens microflow channels in the drainage area
cal solution. of the stimulated well. Here, we present the simulation of a hy-
draulic-fracturing problem as an example case to illustrate the
capability of our hybrid fracture model to capture such a complex
Verification for Linear Flow With Geomechanics. Wu and fracture network in these reservoirs. Three different fracture mod-
Pruess (2000) presented an analytical method for analyzing the els (as shown in Fig. 12) are built, and their flow behavior is com-
nonlinear coupled rock-permeability-variation/fluid-flow problem. pared. The first one considers that there is no natural-fracture-
Approximate analytical solutions for 1D linear and radial flow are active area, and the whole formation is single-porosity shales with
obtained by an integral method, which is widely used in the study low permeability. In the second model, we assume that only the
of steady and unsteady heat-conduction problems. The accuracy natural fractures within the SRV near the hydraulic fractures are
of integral solutions is generally acceptable for engineering appli- active and the rest of the natural fractures outside the SRV remain
cations. When applied to fluid-flow problems in porous media, the inactive. An increase in pore pressure around the hydraulic frac-
integral method consists of assuming a pressure profile in the ture causes a significant reduction in the effective stresses, poten-
pressure-disturbance zone and determining the coefficients of the tially reopening the existing healed natural fractures or creating
profile by making use of the integral mass-balance equation. new fractures. As a result, the permeability near the well of the
The parameters, as shown in Table 1, are used to evaluate reservoir is significantly improved. This effect would help
both the numerical solution and the integral solution. A compari- increase the well productivity in the initial production. The third
son of injection pressures from integral and numerical solutions is fracture model is that all the formation is naturally fractured.
shown in Fig. 10. The agreement between the two solutions is To simulate the performance of this system with our model,
excellent for the entire transient period. hydraulic fractures are represented by the discrete-fracture model
and an active, naturally fractured reservoir area is described by
Model Application the multicontinuum-fracture model, whereas a nonactive-natural-
In the following model-application examples, we are concerned fracture reservoir area is represented by the single-porosity model.
with gas flow toward one horizontal well and a 10-stage hydrau- The basic parameter set for the simulation and discussion is sum-
lic-fracture system in an extremely tight, uniformly porous and/or marized in Table 2, which are chosen field data.
fractured reservoir (Fig. 11). The reservoir formation is at liquid/ We first compare the gas-production behavior for these three
gas, two-phase condition; however, the liquid saturation is set at fracture models. Then, on the basis of the second fracture model
residual values as an immobile phase. This is a single-phase gas- (i.e., reactivated natural fractures only in SRV), we analyze the
flow problem and is modeled by the two-phase-flow reservoir cumulative-gas-production curves with the Klinkenberg, geome-
simulator. The immobile liquid flow is controlled by liquid rela- chanics, and adsorption/desorption effects.
tive permeability curves.
We demonstrate the application of the proposed mathematical
model for modeling gas production from a producer with 10-stage
hydraulic fracturing in a shale gas reservoir. The stress alteration
induced by hydraulic fracturing may activate existing natural frac-

Fig. 12—Three different fracture models: From left to right are


no-natural-fracture model, SRV model, and all-formation-natu-
Fig. 11—Horizontal and multistaged hydraulic-fracture model. rally-fractured model.

852 October 2014 SPE Journal

ID: jaganm Time: 16:29 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 853 Total Pages: 13

TABLE 2—DATA USED FOR THE CASE STUDIES

Reservoir length, Dx, ft 5,500 Hydraulic-fracture permeability, khf , md 1105


Reservoir width, Dy, ft 2,000 Natural-fracture porosity, Unf 0.001
Formation thickness, Dz, ft 250 Natural-fracture total compressibility, cnf , psi1 2.5104
Reservoir depth, h, ft 5,800 Natural-fracture permeability, knf , md 1,600
Reservoir temperature, T, oF 200 Matrix total compressibility, ctm , psi1 2.5104
Initial reservoir pressure, Pi , psi 3,800 Matrix permeability, km , md 3.2105
Horizontal well length, Lh , ft 4,800 Matrix porosity, Um 0.05
Constant flowing bottomhole pressure, Pwf , psi 1,000 Viscosity, l, cp 0.0184
Hydraulic-fracture number 10 Langmuir’s volume, VL , scf/ton 77.56
Distance between hydraulic fractures, 2ye , ft 500 Langmuir’s pressure, PL , psi 2,285.7
Hydraulic-fracture porosity, Uhf 0.5 Non-Darcy-flow constant, b, ft1 1.29106
Hydraulic-fracture total compressibility, chf , psi1 2.5104
Hydraulic-fracture half-length, Xf , ft 250

8000 105 md, and the initial reservoir pressure is 3,800 psi. With this
Cumulative Production, MMscf

permeability value and under higher pressure, the Klinkenberg


7000 effect will not have an obvious influence on gas-flow permeability
6000 on the basis of the estimation in Fig. 3. However, the constant bot-
tomhole production pressure is set as 1,000 psi, which is much
5000 smaller than the reservoir initial pressure. When the pressure of
4000 the region near the wellbore and hydraulic fracture decreases
quickly, the Klinkenberg effect becomes important for the flow in
3000 this region. On the basis of Eqs. 12 and 13, the effective perme-
No Fracture ability considering the Klinkenberg effect at initial pressure
2000
SRV (3,800 psi) is 3:69  105 md, whereas that at bottomhole pres-
1000 Total Fracture
sure (1,000 psi) is 5:0  105 md. Our simulation result in Fig. 15
also shows the influence of the Klinkenberg effect. It leads to
0
approximately a 4% increase to the total gas production.
0 20 40 60 80 100 We studied the non-Darcy flow in the preceding scenario of a
Time, years horizontal well with multistage hydraulic fractures and natural
fractures to see its influence on gas production. The simulation
Fig. 13—Simulated gas-production performance for the three result is shown in Fig. 16 and 17. The difference is observed on
fracture models. the gas cumulative production between the case considering the
non-Darcy flow and the case not considering the non-Darcy flow
Fig. 13 compares the performance of the fractured horizontal in the first 6 years. Not considering the non-Darcy flow inside hy-
well for the three fracture models. The comparison indicates that draulic fractures could lead to an overestimate of approximately
the fracture model makes a difference in well performance. The 5% of cumulative gas production. After that, the difference
contribution from active natural fractures is evident and helps to between cases diminishes until these two curves coincide at
yield higher production rates for a long period. A larger SRV approximately 40 years.
leads to a higher gas-production rate. This simulation result is reasonable with the following analy-
For the second fracture model, pressure distributions at 1 year sis. In Fig. 18, we compare the calculated gas-flow velocities
and 20 years are presented in Fig. 14. from Darcy’s law and the Forchheimer equation for different pres-
Fig. 15 shows the cumulative-production comparison between sure gradients. The parameters of permeability, viscosity, and the
cases with and without the Klinkenberg effect. Here, our simula- non-Darcy-flow factor in this calculation are the same as those in
tor handles the Klinkenberg beta factor not as a constant value, Table 2. When the pressure gradient is less than 1:0  103 psi/ft
but a changing value with matrix permeability and pressure. As or velocity is less than10 ft/D, there is almost no difference
shown in Table 2, the input data of matrix permeability are 3:2  between these two calculations. However, if the pressure gradient

Pg Pg
2.4E+07 2.4E+07
2.2E+07 2.2E+07
2E+07 2E+07
1.8E+07 1.8E+07
1.6E+07 1.6E+07
1.4E+07 1.4E+07
1.2E+07 1.2E+07
1E+07 1E+07
8E+06 8E+06
6E+06 6E+06
4E+06 4E+06

Fig. 14—Pressure distribution at 1 year (left) and 20 years (right) of Fracture Model #2 (unit: Pa).

October 2014 SPE Journal 853

ID: jaganm Time: 16:30 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 854 Total Pages: 13

6000 3000

Cumulative Prodcution, MMscf


Total Prodcution, MMscf 5000 2500

4000 2000

3000
1500
case without Darcy flow
2000
Klinkengberg 1000
case with non-Darcy flow
1000 Klinkenberg
500
0
0 20 40 60 80 100 0
Time, years 0 2 4 6 8
Time, years
Fig. 15—Gas-cumulative-production behavior with the Klinken-
berg effect. Fig. 16—Gas-cumulative-production behavior with non-Darcy
flow in the first 6 years.

6000 1.0E+5
Cumulative Production, MMscf

5000 1.0E+4

1.0E+3
4000

Velocity, ft/D
1.0E+2
3000
1.0E+1
Darcy flow
2000
non-Darcy flow Darcy Velocity
1.0E+0
1000 non-Darcy Velocity
1.0E–1
0
1.0E–2
0 20 40 60 80 100
1.0E–6 1.0E–4 1.0E–2 1.0E+0
Time, years
Pressure Gradient, psi/ft
Fig. 17—Gas cumulative-production behavior with non-Darcy Fig. 18—Darcy and non-Darcy velocities with pressure
flow in 100 years. gradient.

keeps increasing from 1:0  103 psi/ft, the difference will input of the figure data. As shown in Fig. 20, geomechanics/flow
become larger. Fig. 19 shows the calculated average gas flow rate coupling has a large impact on formation permeability, especially
inside hydraulic fractures with production time in the case that for the natural-fracture system. Consider the Muska formation, for
does not consider the non-Darcy flow. For the first 6 years, flow example, when the effective stress increases from 1,600 to 4,800
velocities locate in the range in which the difference between the psia and permeability decreases to 1/20 of its original value. With
Darcy flow and the non-Darcy flow is obvious. After that, flow the gas production, reservoir effective stress increases as pore
velocities move to the area in which the difference is negligible. pressure decreases, leading to the large reduction of cumulative
Fig. 20 shows the simulated-well cumulative production vs. gas production.
time with and without geomechanics effect. The relationship used Figs. 21 and 22 present the results for adsorption analysis with
for describing the effective stress and permeability of the uncon- the numerical model. On the basis of the data in Table 2, we cal-
ventional reservoir is shown in Fig. 1, by use of a table-lookup culate the total gas mass as free gas in the micropores and

1.0E+5 6000
Cumulative Production, MMscf

1.0E+4 Gas Flow Velocity in HF 5000


Velocity in HF, ft/D

1.0E+3
4000
1.0E+2
3000
1.0E+1
2000 case without
1.0E+0 geomechanics
1000 case with
1.0E–1 geomechanics

1.0E–2 0
0 10 20 30 0 10 20 30 40 50 60 70 80 90 100
Time, years Time, years

Fig. 19—Calculated gas-flow velocity with time in hydraulic Fig. 20—Gas-cumulative-production behaviors with geo-
fractures. mechanics.

854 October 2014 SPE Journal

ID: jaganm Time: 16:30 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 855 Total Pages: 13

6000 6000

Cumulative Production, MMscf


Cumulative Production, MMscf
5000 5000

4000 4000 gas production


from free gas
3000 3000 gas production
from adsorption
total production
2000 2000
case without adsorption
1000 1000
case with adsorption

0 0
0 20 40 60 80 100 0 20 40 60 80 100
Time, years Time, years

Fig. 21—Gas-cumulative-production behaviors with adsorption. Fig. 22—Gas-production analysis from free gas and adsorbed
gas.

adsorbed gas at initial condition. The proportion of gas stored in presented at the SPE Gas Production Conference, Fort Worth, Texas,
the pore space is approximately 77%, whereas that stored as 16–18 November. http://dx.doi.org/10.2118/119892-MS.
adsorption is 23%. Then, we compare the cumulative gas produc- Cipolla, C.L. 2009. Modeling Production and Evaluating Fracture Per-
tion with and without considering adsorption. Simulation results formance in Unconventional Gas Reservoirs. J. Pet Tech 61 (9): 84–90
(Fig. 21) show that the estimated gas production will increase (Distinguished Author Series). http://dx.doi.org/10.2118/118536-PA.
with considering adsorption. This difference will become more Cipolla, C.L., Lolon, E.P., Erdle, J.C. et al. 2010. Reservoir Modeling in
and more evident. For the situation considering gas adsorption/de- Shale-Gas Reservoirs. Res Eval & Eng. 13 (4): 638–653. http://
sorption, gas production from the desorption is approximately dx.doi.org/10.2118/125530-PA.
13%, and the produced portion of the free gas consists of 87%, as Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J. et al. 2009. The Rela-
shown in Fig. 22. tionship Between Fracture Complexity, Reservoir Properties, and
Fracture Treatment Design. Paper SPE 115769 presented at the SPE
ATC&E, Denver, Colorado, 21–24 September. http://dx.doi.org/
Summary and Conclusions 10.2118/115769-MS.
This paper discusses a generalized-framework mathematical Davies, J.P. and Davies, D.K. 1999. Stress-Dependent Permeability Char-
model for modeling gas production from unconventional gas res- acterization and Modeling. Paper SPE 56813 presented at the SPE
ervoirs. The model formulation incorporates known nonlinear ATCE, Houston, Texas, 3–6 October. http://dx.doi.org/10.2118/
flow processes, associated with gas production from low-perme- 56813-MS.
ability unconventional reservoirs, including the Klinkenberg, EIA (US Energy Information Administration). 2011. World Shale Gas
non-Darcy-flow, and nonlinear-adsorption effects. The model for- Resources: An Initial Assessment of 14 Regions Outside the United
mulation and numerical scheme are based on a generalized two- States (April).
phase (gas/liquid) -flow model with unstructured grids. Specifi- Ertekin, T., King, G.R., and Schwerer, F.C. 1986. Dynamic Gas Slippage:
cally, a hybrid modeling approach is presented by combining dis- A Unique Dual-Mechanism Approach to the Flow of Gas in Tight For-
crete fracture, multidomain, and multicontinuum concepts for mations. SPE Form Eval 1 (1): 43–52. http://dx.doi.org/10.2118/
handling hydraulic fractures and a fracture network in SRV, dis- 12045-PA.
tributed natural fractures, microfractures as well as porous matrix. Evans, R.D., and Civan, F. 1994. Characterization of Non-Darcy Multi-
We have verified the numerical models against analytical solu- phase Flow in Petroleum Bearing Formation. Final Report for US
tions for the Klinkenberg, non-Darcy-flow, and nonlinear-adsorp- Department of Energy Assistant Secretary for Fossil Energy.
tion effects. Forsyth, P.A., Wu, Y.S., and Pruess, K. 1995. Robust Numerical Methods
As application examples, we present modeling studies with for Saturated-Unsaturated Flow With Dry Initial Conditions in Hetero-
three fracture models for gas production from a 10-stage hydrau- geneous Media. Advance in Water Resources 18: 25–38.
lic-fractured horizontal well, incorporating the Klinkenberg, non-
Freeman, C.M., Moridis, G.J., Ilk, D. et al. 2010. A Numerical Study of
Darcy-flow, and nonlinear-adsorption effects. The model results
Transport and Storage Effects for Tight Gas and Shale Gas Reservoir
show that there is a large impact of various fracture models on
Systems. Paper SPE 131583 presented at the SPE International Oil and
gas-production rates as well as cumulative production.
Gas Conference and Exhibition, Beijing, China, 8–10 June. http://
dx.doi.org/10.2118/131583-MS.
Acknowledgments Freeman, C.M., Moridis, G.J., and Blasingame, T.A. 2009a. A Numerical
Study of Microscale Flow Behavior in Tight Gas and Shale Gas Reser-
This work was supported in part by EMG Research Center and voir Systems. Proceedings of the 2010 TOUGH Symposium, Berkeley,
UNGI of the Petroleum Engineering Department at Colorado California, 14–16 September.
School of Mines; by Foundation CMG; by RIPED of PetroChina
Freeman, C.M., Moridis, G.J., Ilk, D. et al. 2009b. A Numerical Study of
Company; and by IFPEN.
Tight Gas and Shale Gas Reservoir Systems. Paper SPE 124961 pre-
sented at the SPE Annual Technical Conference and Exhibition, New
Orleans, Louisiana, 4–9 October. http://dx.doi.org/10.2118/124961-
References MS.
Blasingame, T.A. 2008. The Characteristic Flow Behavior of Low-Perme- Gao, C., Lee, J.W., Spivey, J.P. et al. 1994. Modeling Multilayer Gas Res-
ability Reservoir Systems. Paper SPE 114168 presented at the SPE ervoirs Including Sorption Effects. Paper SPE 29173 presented at the
Unconventional Reservoirs Conference, Keystone, Colorado, 10–12 SPE Eastern Regional Conference and Exhibition, Charleston, West
February. http://dx.doi.org/10.2118-MS. Virginia, 8–10 November. http://dx.doi.org/10.2118/29173-MS.
Bustin, R.M., Bustin, A.M.M., Cui, X. et al. 2008. Impact of Shale Proper- Kelkar M. and Atiq, M. 2010. Upgrading Method for Tight Gas Reser-
ties on Pore Structure and Storage Characteristics. Paper SPE 119892 voirs. Paper SPE 133301 presented at the SPE Annual Technical

October 2014 SPE Journal 855

ID: jaganm Time: 16:30 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 856 Total Pages: 13

Conference and Exhibition, Florence, Italy, 19–22. http://dx.doi.org/ Wu, Y.S. 2002. Numerical Simulation of Single-Phase and Multiphase
10.2118/133301-MS. Non-Darcy Flow in Porous and Fractured Reservoirs. Transport in Po-
Klinkenberg, L.J. 1941. The Permeability of Porous Media to Liquids and rous Media 49: 209–240.
Gases. In API Drilling and Production Practice, 200–213. Wu, Y.S. and Pruess, K. 2000. Integral Solutions for Transient Fluid Flow
Langmuir, I. 1916. The Constitution and Fundamental Properties of Solids Through a Porous Medium With Pressure-Dependent Permeability.
and Liquids. J. Am. Chem. Soc. 38 (11): 2221–2295. International J. Rock Mechanics and Mining Sci. 37: 51–61.
Wu, Y.S., Pruess, K., and Persoff, P. 1998. Gas Flow in Porous Media
Leahy-Dios, A., Das, M., Agarwal, A. et al. 2011. Modeling of Transport
With Klinkenberg Effects. Transport in Porous Media 32: 117–137.
Phenomena and Multicomponent Sorption for Shale Gas and Coalbed
Wu, Y.S., Rutqvist, J., Karasaki, K. et al. 2008. A Mathematic Model for
Methane in an Unstructured Grid Simulator. Paper SPE 147352 pre-
Rock Deformation Effect of Flow in Porous and Fractured Reservoirs.
sented at the SPE Annual Technical Conference, Denver, Colorado, 30
Paper ARMA-08-142 presented at the 42nd US Rock Mechanics Sym-
Octomber–2 November. http://dx.doi.org/10.2118/147352-MS.
posium and 2nd US–Canada Rock Mechanics Symposium, San Fran-
Leverett, M.C. 1941. Capillary Behavior in Porous Media. Trans AIME cisco, California, 29 June–2 July.
142: 341–358. Wu, Y.S. and Wang, C. 2012. Transient Pressure Testing Analysis of Gas
Li, J., Du, C.M., and Zhang, X. 2011. Critical Evaluation of Shale Gas Wells in Unconventional Reservoirs. Paper SPE-SAS-312 presented at
Reservoir Simulation Approaches: Single-Porosity and Dual-Porosity the SPE Annual Technical Symposium and Exhibition (ATS&E), Kho-
Modeling. Paper presented at the SPE Middle East Unconventional bar, Saudi Arabia, 8–11 April.
Gas Conference and Exhibition, Muscal, Oman, 31 January–2 Wu, Y.S., Wang, C., Li, J. et al. 2012. Transient Gas Flow in Unconventional
February. Gas Reservoirs. Paper SPE 154448 presented at the EAGE Annual Con-
Mengal, S.A. and Wattenbarger, R.A. 2011. Accounting for Adsorbed Gas ference and & Exhibition Incorporating SPE Europec, Copenhagen, Den-
in Shale Gas Reservoirs. Paper SPE 141085 presented at the SPE Mid- mark, 4–7 June. http://dx.doi.org/10.2118/154448-MS.
dle East Oil and Gas Show and Conference, Manama, Bahrain, 25–28
September. http://dx.doi.org/10.2118/141085-MS. Appendix A
Moridis, G.J., Blasingame, T.A., and Freeman, C.M. 2010. Analysis of Here, we derive the analytical solution for gas flow with adsorp-
Mechanisms of Flow in Fractured Tight-Gas and Shale-Gas Reser- tion/desorption. If the system is isothermal, the ideal-gas law
voirs. Paper SPE 139250 presented at the SPE Latin American and applies, and the gravity effect is negligible, then gas flow in porous
Caribbean Petroleum Engineering Conference, Lima, Peru, 1–3 media with adsorption is described by the following equations:
December. http://dx.doi.org/10.2118/139250-MS.
Ozkan, E., Raghavan, R., and Apaydin, O.G. 2010. Modeling of Fluid
@ð/q þ mg Þ
r  ðqvÞ ¼  ; . . . . . . . . . . . . . . . . . . ðA-1Þ
Transfer From Shale Matrix to Fracture Network. Paper SPE 134830 @t
presented at the SPE ATC&E, Florence, Italy, 19–22 September. where q is the gas density; v is the gas-flow velocity; / is the po-
http://dx.doi.org/10.2118/134830-MS. rous-media porosity; mg is the adsorbed gas mass in a unit forma-
Pruess, K. 1983. GMINC—A Mesh Generator for Flow Simulations in tion volume at a given pressure; and t is the time.
Fractured Reservoirs, Report LBL-15227, Berkeley, California: Law- According to the ideal-gas law,
rence Berkeley National Laboratory.
Pruess, K. and Narasimhan, T.N. 1985. A Practical Method for Modeling PV ¼ nRT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-2Þ
Fluid and Heat Flow in Fractured Porous Media. Soc. Pet. Eng. J. 25:
14–26. and
Pruess, K., Oldenburg, C., and Moridis, G. 1999. TOUGH2 User’s Guide, M
q¼ P ¼ bP; . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-3Þ
Version 2.0, Report LBNL-43134, Berkeley, California: Lawrence RT
Berkeley National Laboratory.
where M is gas molecular weight; R is the universal gas constant;
Rubin, B. 2010. Accurate Simulation of Non-Darcy Flow in Stimulated M
Fractured Shale Reservoirs. Paper SPE 132093 presented at the SPE b is a coefficient, for simplicity, defined as b ¼ ; and T is the
Western Regional Meeting, Anaheim, California, 27–29 May. http://
RT
system temperature.
dx.doi.org/10.2118/132093-MS. From Darcy’s law and the Langmuir isotherm (Eqs. 2 and 3),
Rutqvist, J.Y.,Wu, S., Tsang, C.F. et al. 2002. A Modeling Approach for
Analysis of Coupled Multiphase Fluid Flow, Heat Transfer, and Defor- k
v ¼  rP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-4Þ
mation in Fractured Porous Rock. International J. Rock Mechanics l
and Mining Sci. 39: 429–442.
and
Silin, D. and Kneafsey, T. 2011. Gas Shale: From Nanometer-Scale Obser-
vations to Well Modeling. Paper CSUG/SPE 149489 presented at the P P
mg ¼ qR qg VE ¼ qR qg VL ¼ Va ; . . . . . ðA-5Þ
Canadian Unconventional Resources Conference, Calgary, Alberta, P þ PL P þ PL
Canada, 15–17 November. http://dx.doi.org/10.2118/149489-MS.
where qR is rock bulk density; qg is gas density at standard condi-
Soeder, D.J. 1988. Porosity and Permeability of Eastern Devonian Gas
tion; VE is the adsorption isotherm function for gas content; VL is
Shale. SPE Form Eval 3 (1): 116–124. http://dx.doi.org/10.2118/
the Langmuir’s volume in scf/ton; and PL is Langmuir’s pressure.
15213-PA.
a is a coefficient, for simplicity, defined as a ¼ qR qg VL .
Wang, F.P., Reed, R.M., Jackson, J.A. et al. 2009. Pore Networks and By substituting Eqs. A-4 and A-5 into Eq. A-1, we obtain
Fluid Flow in Gas Shales. Paper SPE 124253 presented at the SPE An-  
nual Technical Conference and Exhibition, New Orleans, Louisiana, P
  @
4–7 October. http://dx.doi.org/10.2118/124253-MS. k @P P þ PL
r  b PrP ¼ /b þa : . . . . . . . .ðA-6Þ
Winterfeld, P.H. and Wu, Y.S. 2011. Parallel Simulation of CO2 Seques- l @t @t
tration With Rock Deformation in Saline Aquifers. Paper SPE 141514
prepared for presentation at the SPE Reservoir Simulation Symposium, In radial coordinates,
 
The Woodlands, Texas, 21–23 February. http://dx.doi.org/10.2118/ P
 2
 @
141514-MS. 1@ @P 2/l @P 2al P þ PL @P
r ¼ þ ; . . . . ðA-7Þ
Winterfeld, P.H. and Wu, Y.S. 2012. A Novel Fully Coupled Geomechan- r @r @r k @t bk @P @t
ical Model for CO2 Sequestration in Fractured and Porous Brine Aqui-
fers. Paper presented at the XIX International Conference on Water   " #
1@ @P2 2/l 2al PL @P
Resources CMWR 2012, University of Illinois at Urbana-Champaign, r ¼ þ ; . . . . . . ðA-8Þ
r @r @r k bk ðP þ PL Þ2 @t
Illinois, 17–22 June.

856 October 2014 SPE Journal

ID: jaganm Time: 16:30 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


J163609 DOI: 10.2118/163609-PA Date: 9-October-14 Stage: Page: 857 Total Pages: 13

and earned BS (Eqv.) and MS degrees in petroleum engineering in


China, and MS and PhD degrees in reservoir engineering from
  " #
1@ @P2 /l al PL @P2 the University of California at Berkeley. At CSM, he is teaching
r ¼ þ 2
: . . . . . . ðA-9Þ petroleum reservoir engineering courses, supervising graduate
r @r @r Pk Pbk ðP þ PL Þ @t students, and conducting research in the areas of multiphase
fluid and heat flow in porous media, enhanced oil recovery
Eq. 9 becomes (EOR), CO2 geosequestration and CO2 EOR, reservoir simula-
tion, enhanced geothermal systems, geomechanics coupling
1 @ @P2 1 @P2 and rock-deformation effects, and unconventional hydrocar-
ðr Þ¼ ; . . . . . . . . . . . . . . . . . . . . . ðA-10Þ
r @r @r A @t bon reservoirs. Previously, Wu was a staff scientist with LBNL for
14 years (1995 through 2008).
where we define the coefficient
Jianfang Li is a senior research engineer at Research Institute
1 /l al PL of Petroleum Exploration and Development (RIPED), Petro-
¼ þ : . . . . . . . . . . . . . . . . . . ðA-11Þ China. She earned BS and MS degrees in applied mathemat-
A Pk Pbk ðP þ PL Þ2 ics from Beijing University of Aeronautics and Astronautics and
a PhD degree in reservoir engineering from the graduate
We propose to use a history-dependent, constant, averaged school of RIPED. Li’s research interests include optimization,
pressure within the pressure-changed domain (Wu et al. 1998), algorithms for solving linear algebraic equations, reservoir simula-
X tion (carbonate-reservoir simulation, chemical-flooding reservoir
Vj Pj simulation, natural-gas-reservoir simulation, and unconventional-
P X ; . . . . . . . . . . . . . . . . . . . . . . . . . . ðA-12Þ hydrocarbon-reservoir simulation).
Vj
Didier-Yu Ding is a senior research engineer at IFP Energies
where Vj is a controlled volume at the geometric center of which Nouvelles in France. His research interests include numerical
the pressure was Pj at the immediately preceding modeling and optimization, reservoir simulation and charac-
X time when the terization, complex wells, and near-well flow. Ding holds a BS
solution was calculated. The summation, Vj , is performed degree in mathematics from Peking University in China, and
over all Vj in which pressure increases (or decreases) occurred at MS and PhD degrees in applied mathematics from University
the preceding time value. Pj is always evaluated analytically at de Paris, France.
point j, on the basis of the previous estimated, constant Cong Wang is a PhD candidate in petroleum engineering at
diffusivity. CSM. He earned a BS degree from Peking University and an
The well boundary proposed as a line source/sink well is MS degree from the Petroleum Engineering Department, CSM.
Wang’s Master’s degree study was on the development of an
pkhrb @P2 unconventional reservoir simulator and its application in tran-
lim ¼ Qm : . . . . . . . . . . . . . . . . . . . . . ðA-13Þ sient-pressure analysis and fracture characterization. He has
x!0 l @r
made several conference presentations at SPE events, and his
Then, we could get a transient-pressure solution for gas flow current research involves fully coupling geomechanics and
with adsorption/desorption, fluid flow in unconventional reservoirs.
  Yuan Di is an associate professor in the Department of Energy
lQm r2 and Resources Engineering, College of Engineering, Peking
P2 ðr; tÞ ¼P2i  Ei  :             ðA-14Þ University. He earned a BS degree from Xi’an Jiaotong Univer-
2pkhrb 4At
sity, an MS degree from Xi’an University of Architecture and
Technology, and a PhD degree from Tongji University, Shang-
Yu-Shu Wu is a professor and Foundation CMG Research Chair hai—all in civil engineering. Di’s research interests include the
in Petroleum Engineering at the Colorado School of Mines numerical simulation of multiphase flow in porous media,
(CSM). He is also a guest scientist at the Earth Sciences Division wave propagation in porous media, and the analysis of bore-
of the Lawrence Berkeley National Laboratory (LBNL). Wu hole stability.

October 2014 SPE Journal 857

ID: jaganm Time: 16:30 I Path: S:/3B2/J###/Vol00000/140016/APPFile/SA-J###140016


Reservoir Modeling in Shale-Gas Reservoirs
C.L. Cipolla and E.P. Lolon, StrataGen Engineering; and J.C. Erdle and B. Rubin, Computer Modelling Group

Summary many shale reservoirs was made possible through improved stimu-
The exploitation of unconventional gas reservoirs has become an lation techniques and horizontal drilling.
ever increasing component of the North American gas supply. The economic viability of many unconventional gas develop-
The economic viability of many unconventional gas developments ments hinges on effective stimulation of extremely low-permeability
hinges on effective stimulation of extremely low-permeability rock rock, typically 10 to 100 nanodarcies (10−6 md). In most cases,
by creating very complex fracture networks that connect huge economic production is possible only if a very complex, highly
reservoir surface area to the wellbore. In addition, gas desorption nonlinear fracture network can be created that effectively connects
may be a significant component of overall gas recovery in many a huge reservoir surface area to the wellbore. Many conventional
shale-gas reservoirs. The widespread application of microseismic hydraulic-fracture treatments use high-viscosity fluids to reduce
(MS) mapping has significantly improved our understanding of fracture complexity and promote planar fractures, allowing the
hydraulic fracture growth in unconventional gas reservoirs (pri- placement of high concentrations of large proppant. However,
marily shale) and has led to better stimulation designs. However, stimulation treatments in shale reservoirs use large volumes of
the overall effectiveness of stimulation treatments is difficult to low-viscosity fluid (water) to promote fracture complexity and
determine from MS mapping because the location of proppant place very low concentrations of small proppant—an approach
and the distribution of conductivity in the fracture network cannot that is completely different from conventional hydraulic fracturing.
be measured (and are critical parameters that control well perfor- Hydraulic-fracture treatment stages in horizontal wells can require
mance). Therefore, it is important to develop reservoir-modeling over a million gallons of water and up to 1 million pounds of prop-
approaches that properly characterize fluid flow in and the prop- pant and are pumped at rates of 75–150 bbl/min. Because of the
erties of a complex fracture network, tight matrix, and primary complexity of fracture growth in many shale-gas reservoirs, it is not
hydraulic fracture (if present) to evaluate well performance and possible to accurately predict the distribution of proppant within
understand critical parameters that affect gas recovery. the network and the conductivity of the network fractures is dif-
This paper illustrates the impact of gas desorption on produc- ficult to predict. Previous work has characterized proppant distribu-
tion profile and ultimate gas recovery in shale reservoirs, showing tion within complex fracture networks using two limiting cases:
that in some shale-gas reservoirs desorption may be a minor com- either the proppant is evenly distributed throughout the complex
ponent of gas recovery. In addition, the paper details the impact of network or the proppant is confined to a primary or main fracture
changing closure stress distribution in the fracture network on well (Cipolla et al. 2008a). This previous work showed that for prop-
productivity and gas recovery. In shale-gas reservoirs with lower pant volumes that are typical for many shale-gas hydraulic-fracture
Young’s modulus rock, stress-dependent network-fracture conduc- treatments, if the proppant is evenly distributed throughout the
tivity may reduce ultimate gas recovery significantly. The paper complex fracture network, proppant concentrations (lbm/ft2) are
includes an example that contrasts the application of numerical likely to be too low to materially affect network-fracture conductiv-
reservoir simulation and advanced decline-curve analyses to illus- ity and the network fractures will essentially behave as unpropped
trate issues associated with conventional production-data-analysis fractures. If the proppant is confined to a primary or main hydraulic
techniques when applied to unconventional reservoirs. fracture and is not transported into the network fractures, then the
Selected examples from the Barnett shale are included that relative conductivity of the primary fracture should be high, but the
incorporate MS fracture mapping and production data to illustrate fracture network will be unpropped. Therefore, unpropped-fracture
the application of production modeling to evaluate well perfor- conductivity may play an important role in shale-gas stimulation
mance in unconventional gas reservoirs. This paper highlights performance. Network-fracture conductivity in many shale-gas
production modeling and analysis techniques that aid in evaluat- reservoirs is estimated by modeling well performance. Combining
ing stimulation and completion strategies in unconventional gas detailed reservoir modeling with MS-mapping data that define the
reservoirs. size of the fracture network and laboratory cores tests that provide
estimates of matrix permeability and propped- and unpropped-
Introduction fracture conductivity can result in a much better understanding of
stimulation performance.
Gas shales are organic-rich shale formations and are apparently
Gas flow from ultralow-permeability rock through the complex
both the source rock and the reservoir. The gas is stored in the
fracture network must be modeled to evaluate stimulation designs
limited pore space of these rocks and a sizable fraction of the gas
and completion strategies properly. Therefore, the complex frac-
in place may be adsorbed on the organic material. The natural-
ture network and primary hydraulic fracture (if present) must be
gas resource potential for gas shales in the USA is estimated to
discretely characterized in our reservoir-simulation models. MS
be from 500 to 1,000 Tcf (Arthur et al. 2008). Typical shale-gas
fracture mapping provides a measurement of the overall reservoir
reservoirs exhibit a net thickness of 50 to 600 ft, porosity of 2 to
volume that is stimulated [referred to as stimulated reservoir vol-
8%, and total organic carbon of 1 to 14% and are found at depths
ume (SRV)], and special core analysis can provide measurements
ranging from 1,000 to 13,000 ft. The success of the Barnett shale
of matrix permeability. With these two parameters, reservoir-simu-
has illustrated that gas can be economically produced from rock
lation models can be used to estimate the spacing of the network
that was previously thought to be source rock and/or caprock, not
fractures (complexity) and their conductivity. Simulated produc-
reservoir rock. This revelation has led to the development of many
tion profiles can then be compared to actual well performance to
other shale-gas reservoirs, including the Woodford, Fayetteville,
evaluate the effectiveness of stimulation designs and completion
Marcellus, and the Haynesville (Fig. 1). In addition to increasing
strategies and diagnose the relative conductivity of the primary
natural-gas prices (until recently), the economic development of
fracture (if present). This paper builds upon and extends previous
shale-gas reservoir-modeling work (Mayerhofer et al. 2006, 2010;
Copyright © 2010 Society of Petroleum Engineers
Warpinski et al. 2008; Cipolla et al. 2008a, 2009b), focusing on
This paper (SPE 125530) was accepted for presentation at the SPE Eastern Regional the impact of gas desorption and stress-sensitive network-fracture
Meeting, Charleston, West Virginia, USA, 23–25 September 2009 and revised for publication.
Original manuscript received for review 21 July 2009. Revised manuscript received
conductivity on well performance. In addition, this work discusses
for review 29 December 2009. Paper peer approved 1 March 2010. various shale-gas reservoir-simulation approaches, detailing an

638 August 2010 SPE Reservoir Evaluation & Engineering


Fig. 1—Lower 48 states shale-gas plays.

approach that appears best suited for most shale-gas reservoirs. stimulation fluid from the treatment well, confirming that large
The highlights of the previous work are summarized to provide volumes of water are transversing the complex fracture network.
background to support the work presented in this paper. The volume of reservoir rock that is contacted by the stimulation
treatment (SRV), as evidenced by the MS activity, has a direct
Background impact on well performance—with larger SRVs resulting in better
The initial MS-mapping work in the Barnett shale revealed that wells (Mayerhofer et al. 2010). This early MS-mapping work led
fracture growth can be surprisingly complex in some environments. to a much better understanding of fracture growth in the Barnett
Fig. 2 shows the MS-mapping data for a typical vertical well in shale and was instrumental in many improvements in stimulation
the Barnett shale (Fisher et al. 2002). Stimulation treatments in and completion strategies, including horizontal drilling.
vertical wells consist of approximately 750,000 gal of slickwater Fig. 3 shows the MS-mapping data for a horizontal Barnett
and 150,000 lbm of proppant pumped at concentrations of 0.1 to shale well completed with an uncemented liner (five sets of per-
0.25 ppa and rates of 50–70 bbl/min. Early work showed that very forations) and stimulated using a single-stage hydraulic-fracturing
large fracture networks were created in the Barnett shale, typically treatment consisting of 4.2 million gal of slickwater and 853,000
in excess of 2,500 ft long and 1,000 ft wide, covering the Barnett lbm of proppant pumped at 130 bbl/min [Fisher et al. 2004]. In this
reservoir thickness (≈300 ft). In Fig. 2, the fracture network is case, the single-stage hydraulic-fracture treatment covered most of
more than 3,000 ft long and approximately 1,000 ft wide, extend- the 2,400 ft long lateral, but there are areas with little MS activity
ing more than 70 acres and contacting 900 million ft3 of reservoir that may not be stimulated effectively. The stimulation treatment
rock. Fig. 2 also shows that five offset wells were “killed” by appears to have effectively covered an area of approximately

1500 2000.00

Wells killed by stimulation fluid 1600.00

1000 1200.00

800.00
500 Fairway “Length”
“Length”
400.00
Observation Well Fairway“Width”
Fairway “Width”
0.00
South-North (ft)

0
South-North (ft)

–400.00

–500 –800.00
Treatment Well
–1200.00
–1000
–1600.00

–2000.00
–1500
–2400.00

–2000 –2800.00

Wells killed by stimulation fluid –3200.00


–2500
–3600.00

–4000.00
–3000
–2800

–2400

–2000

–1600

–1200

–800

–400

400

800

1200

1600

2000

2400

2800

–1000 –500 0 500 1000 1500 2000 2500 3000


West-East (ft) West-East (ft)

Fig. 2—MS-mapping data from a vertical Barnett shale well Fig. 3—MS-mapping data for a horizontal Barnett shale well,
(Fisher et al. 2002). uncemented liner, single-stage treatment (Fisher et al. 2004).

August 2010 SPE Reservoir Evaluation & Engineering 639


2,200×2,000 ft (square), or 100 acres, not that much larger than matrix. A more detailed discussion of shale-gas reservoir modeling
the vertical-well treatment shown in Fig. 2. This led operators to is provided in Appendix A.
cased-and-cemented horizontal completions in which multistage
stimulation treatments can be pumped, resulting in larger SRVs Gas Desorption and Stress-Dependent
and possibly more-complex fracture networks—with horizontal Fracture Permeability
completions yielding more than three times the estimated ultimate Previous reservoir-modeling work has largely neglected the effect
recovery (EUR) of vertical wells (Frantz et al. 2005). Although of gas desorption and stress-dependent fracture permeability on
MS mapping provided invaluable insights into fracture growth and well productivity and gas recovery. A series of reservoir simula-
the impact of SRV on well performance, there was still a need to tions was performed using reservoir properties typical for the
better understand the mechanisms that controlled production in the Barnett and Marcellus shales to evaluate the likely impact of
Barnett shale (and other shale reservoirs) to improve completion gas desorption and stress-dependent fracture permeability in two
strategies and stimulation designs, which required detailed numeri- important shale developments in the USA (Fig. 1). The simulations
cal reservoir modeling. focus on horizontal completions because the majority of shale-gas
In addition to confirming the relationship between SRV and reservoirs are being developed using cased-and-cemented hori-
well performance, reservoir-modeling work has shown that the zontal wells with multiple fracture treatment stages. The length of
complexity and conductivity of the fracture network are key the horizontal section is 3,000 ft. The reservoir model consists of
components that control well productivity in shale-gas reservoirs main, or primary, fractures and network, or secondary, fractures.
(Mayerhofer et al. 2006; Cipolla et al. 2008a). In addition, the The main (or primary) fracture representing the propped fracture
location of the proppant within the complex fracture network and is connected to the wellbore and is oriented perpendicular to the
the conductivity of the primary or main fracture (if present) play direction of the horizontal wellbore. The primary-fracture spacing
a significant role in well performance (Cipolla et al. 2009a). This is controlled by the number of fracture-treatment stages along the
previous work showed that more-complex fracture networks (i.e., lateral, with more stages resulting in closer spacing of the primary
smaller blocks or more-dense fracture spacing) result in higher fractures. The secondary (or network) fractures are assumed to be
production rates. Reservoir-simulation history matching has indi- unpropped and have no direct communication with the wellbore
cated that network-fracture conductivity may be relatively low and are placed in the perpendicular and parallel directions to the
in the Barnett shale, ranging from 0.5 to 5.0 md-ft, limiting gas primary fracture (assumed square in dimension). The spacing
production in many cases (Mayerhofer et al. 2006; Cipolla et al. of the secondary fractures (size of the network blocks) is varied
2008a). In addition, the presence of a high-relative-conductivity within a range that appears consistent with actual well performance
primary fracture will improve production significantly when net- and estimates of matrix permeability and fracture conductivity
work-fracture conductivity is low; however, work to date suggests (Fredd et al. 2001; Mayerhofer et al. 2006; Cipolla et al. 2008a,
that high-conductivity primary fractures may not be present in the 2009c). It is assumed that gas flow occurs into the wellbore
Barnett shale (Cipolla et al. 2009c). With a better understanding through the main fractures only. A wide range of reservoir and
of the mechanisms that control well productivity, operators have fracture parameters was evaluated, and only a limited number of
focused on completion techniques and stimulation designs that simulations are presented to illustrate the results of this work. All
maximize network complexity, increase SRV, and improve fracture simulations are single phase (dry gas).
conductivity such as more fracture-treatment stages, larger volume
and higher injection rate treatments, diversion to increase fracture Barnett Shale: Impact of Gas Desorption. Table 1 shows the
complexity, larger proppant volumes and higher-strength and/or main reservoir parameters used for Barnett shale reservoir simula-
lower-density proppants to increase fracture conductivity. In addi- tions. Pore pressure is 3,800 psi or 0.54 psi/ft. The reservoir size
tion, operators are experimenting with simultaneous or alternating is 3,400×2,400 ft, with a fracture-network size of 2,000×3,400 ft
stimulations in offsetting wells to increase fracture complexity and (SRV=2,040×106 ft3). Fig. 4 shows a graph of gas content (scf/ton)
improve gas recovery. vs. pressure for the gas porosity and adsorbed gas used for the
Shale-Gas Reservoir Simulation Barnett example (Montgomery 2004). The gas in the matrix poros-
ity is shown as a nearly straight line (circles) while the desorption
Given the complex nature of hydraulic-fracture growth and the very isotherm is represented by a curved line (squares). Both data sets
low permeability of the matrix rock in many shale-gas reservoirs— add together to form the total gas content. In this case, 41% of
in combination with the predominance of horizontal completions— the total gas is adsorbed. At higher pressures, more gas is stored
reservoir simulation is commonly the preferred method to predict in the matrix porosity and the gas porosity will contribute more
and evaluate well performance. However, semianalytical solutions to production than desorption because of the difference in curve
for hydraulically fractured horizontal wells in fractured reservoirs shapes (straight line vs. curved line). Although there appears to
have been published (Medeiros et al. 2008). Analytical solutions be a considerable amount of adsorbed gas, it may be difficult to
for fluid flow in naturally fractured reservoirs were published by capture this gas when reservoir permeability is low.
Warren and Root (1963) and Kazemi (1969); however, analytical Figs. 5 and 6 show the impact of gas desorption on cumulative
solutions to fluid flow in naturally fractured reservoirs typically gas production for a typical Barnett horizontal completion. The
cannot capture the very long transient behavior in the matrix blocks
exhibited by shale-gas reservoirs (caused by the very low perme-
ability of the shale matrix). These “pseudosteady-state” solutions
to fluid flow in naturally fractured reservoirs have been adapted TABLE 1—BARNETT SHALE RESERVOIR PROPERTIES
to numerical reservoir-simulation models to improve run time, but
Depth (ft) 7,000
again, they lack the ability to model transient flow in the matrix
blocks. Techniques have been developed to model the transient Reservoir pressure (psi) 3,800
behavior of matrix blocks in reservoir simulators, but many of Pore pressure gradient (psi/ft) 0.54
these techniques can still rely on analytical approximations that Closure pressure (psi) 5,000
are used within the numerical model to reduce run time. The most Closure pressure gradient (psi/ft) 0.71
rigorous method to model shale-gas reservoirs is to grid the entire
Net pay (ft) 300
reservoir discretely, including the network fractures, hydraulic
fracture, matrix blocks, and unstimulated areas—but this increases Porosity (% ) 3
computational time. However, with the continual advances in Matrix permeability (md) 1 E 5; 1 E 4
computing power, much-more-complex numerical models can be Water saturation (%) 30
used efficiently. The reservoir simulations in this paper were per- Reservoir temperature (°F) 180
formed using a detailed numerical grid that rigorously represents
the complex fracture network, primary fracture, and tight shale FBHP (psi) 1,000

640 August 2010 SPE Reservoir Evaluation & Engineering


Fig. 4—Example of Barnett gas content, adsorbed and free.

Fig. 5—Impact of desorbed gas, Barnett shale (k = 10−4 md,


uniform conductivity network = 2 md-ft).

simulations assume a uniform network-fracture conductivity of


2 md-ft. The production for two cased-and-cemented Barnett hori- the network-fracture spacing is smaller (50-ft square blocks), the
zontal wells is shown for comparison, Well A and Well B. These desorbed gas results in roughly 15% more production after 30
wells were selected because MS-mapping data were available to years (for both reservoir permeability cases). Thus, the impact of
determine the SRV. The SRV, lateral length, and primary-fracture desorption increases when the network-fracture spacing is smaller,
spacing were similar to many of the simulation cases. Well A has with most of the additional gas produced later in the well life. Other
a lateral section of 2,600 ft. The fracture treatments in this well studies have also reported that desorption does not play a major
consisted of four stages (one perforation cluster per stage, with role in the production profile or gas recovery in the Barnett shale
cluster spacing of 700 ft). Therefore, the primary- or main-frac- (Frantz et al. 2005).
ture spacing is 700 ft. The total amount of proppant pumped was The production profile for Well B follows the simulation
670,000 lbm (40/70 sand) with 120,000 bbl of fluid. The MS map- results for the larger network block size (Dx = 600 ft) and matrix
ping of this well showed an SRV of 1,880×106 ft3, which is close permeability of 10−4 md relatively closely, indicating that fracture
to the simulated SRV of 2,000×106 ft3. Well B has a lateral length complexity in Well B may be low. The production profile for Well
similar to that of Well A (2,600 ft).The fracture treatment for Well A follows a similar but somewhat higher trend, possibly indicating
B consisted of 830,000 lbm of 40/70 sand and 117,000 bbl of fluid a more-complex fracture network. It should be emphasized that
pumped in two stages with six perforation clusters, three per stage the production comparisons for Wells A and B are qualitative only
(spaced 500 ft apart). The primary-fracture spacing for Well B is and are not intended to be detailed “history matches.” However,
500 ft. The MS mapping of Well B indicated an SRV of 2,017×106 they provide actual production data from Barnett horizontal wells
ft3, which is also very close to the simulation model. with SRV and primary-fracture spacing similar to that used in
In Fig. 5, (k = 10−4 md) and Fig. 6 (k = 10−5 md), the primary- the reservoir simulations and could be considered “approximate”
fracture spacing is 600 ft for the simulated data (solid and dashed history matches.
lines). The secondary- or network-fracture spacing was varied from Fig. 7 shows pressure distribution after 1, 5, and 15 years for
50 to 600 ft. For the larger network-spacing case (600 ft apart), the two different reservoir-permeability cases. Because of the sym-
the desorbed gas contributed approximately 8.5% of the total gas metry of the reservoir area investigated, only one-half of the total
production after 30 years when the reservoir permeability was 10−4 reservoir size is simulated. The primary fractures extend 1,000
md, and only 6.9% after 30 years when the reservoir permeability ft from the wellbore (half-length). The fracture conductivity is
was 10−5 md. Figs. 5 and 6 show that the desorbed-gas contribution 2 md-ft for both the primary and network fractures (uniform con-
is relatively small within the first 5 years of production and, thus, ductivity). The pressure distribution includes the effect of desorp-
not likely to have a material impact on production economics. If tion. Dark red indicates initial reservoir pressure, while dark blue
indicates flowing bottomhole pressure (FBHP) (1,000 psi). The
left side of the graph shows the pressure distribution for a matrix
permeability of 10−5 md, primary fracture spacing of 300 ft, and
a network block size (Dx) of 300 ft. When matrix permeability is
very low and network-block size is large, pressures in the matrix
blocks remain relatively high even after 15 years of production.
The right side of the graph shows the pressure distribution for a
matrix permeability of 10−4 md, primary fracture spacing of 100 ft,
and network-block size of 50 ft, showing that when network-block
size is small (i.e., more-complex fracture network), the tight matrix
can be effectively drained after 15 years. However, because of the
relatively high FBHP and the Barnett desorption isotherm (Fig. 4),
very little gas is desorbed from the shale even with efficient drain-
age of the tight matrix.
Impact of FBHP. The impact of FBHP is illustrated in Figs. 8
and 9. Figs. 8 and 9 compare the 30-year gas recovery with and
without gas desorption for a FBHP of 500 and 1,000 psi, a typi-
cal range for many Barnett shale wells. The simulations in Fig. 8
assume a matrix permeability of 10−4 md and a network-fracture
spacing of 600 ft, while the simulations in Fig. 9 assume a matrix
Fig. 6—Impact of desorbed gas, Barnett shale (k = 10−5 md, permeability of 10−5 md and a network-fracture spacing of 100 ft.
uniform conductivity network = 2 md-ft). The primary-fracture spacing is 600 ft, and a uniform network

August 2010 SPE Reservoir Evaluation & Engineering 641


k = 1e-5 mD, primary spacing = 300 ft, Dx = 300 ft k = 1e-4 mD, primary spacing = 100 ft, Dx = 50 ft

1 year

5 years

15 years

Fig. 7—Pressure distribution after 1, 5, and 15 years for the Barnett shale example (with desorbed gas).

conductivity of 2 md-ft is used in the simulations in Figs. 8 and 9. result in similar production profiles, but the modeling may indicate
Fig. 9 shows that reducing the FBHP from 1,000 to 500 psi results that Well A has a more complex fracture network compared with
in increases in cumulative gas production of up to 10% after 30 Well B. Non-unique solutions to production history matching
years for the case with 10−4 md matrix permeability (Fig. 8) and are common in shale-gas reservoirs because matrix permeability,
12.5% after 30 years of production for the case with 10−5 md matrix network-fracture spacing, and the conductivity of the network
permeability (Fig. 9). Comparing the results with desorbed-gas and primary fractures are not known exactly. However, laboratory
production (solid lines) to those without desorbed-gas production core tests can be performed to estimate matrix permeability and
(dashed lines) in Figs. 8 and 9 shows that reducing the FBHP from unpropped-fracture conductivity, constraining the solution and
1,000 to 500 psi does not significantly increase the production of allowing significant insights into well performance.
desorbed gas. In addition, reviewing the production profiles for
the first 5 years shows that reducing the FBHP from 1,000 to 500 Barnett Shale: Impact of Stress-Dependent Network-Fracture
psi has very little impact on the early-time production profile, Permeability. The impact of increasing closure stress on propped-
especially in the first 3 years. However, the impact of reducing fracture conductivity has been studied for decades, but little atten-
the FBHP on gas production is greater for more-complex fracture tion has been given to the conductivity of unpropped hydraulically
networks (i.e., smaller blocks). created fractures until recently. There is a likelihood that large
Comparing the production profile from Well A to the simulation portions of the fracture networks created during shale-gas stimula-
results in Fig. 9 shows that Well A follows the production profile tion treatments may be unpropped or ineffectively propped (Fredd
for a matrix permeability of 10−5 md and a network-fracture spacing et al. 2001; Cipolla et al. 2008a). Therefore, unpropped- or partially-
of 100 ft. This is not a unique solution, because other combina- propped-fracture conductivity may play a significant role in shale-
tions of matrix permeability and network-fracture spacing may gas production. The impact of primary- and network-fracture

Fig. 8—Impact of BHFP for the desorbed-gas case (k = 10−4 Fig. 9—Impact of BHFP for the desorbed-gas case (k = 10−5
md, Dx = 600 ft). md, Dx = 100 ft).

642 August 2010 SPE Reservoir Evaluation & Engineering


Fig. 10—Effect of closure stress and Young’s modulus on
unpropped-fracture conductivity.
Fig. 11—Impact of stress-dependent fracture conductivity for
the desorbed-gas case (k = 10−4 md).

conductivity on well performance was presented in previous work


(Cipolla et al. 2008a, 2009a), showing that production in many
shale-gas reservoirs may be limited by insufficient fracture con- was used to calculate the closure stress on the network fractures.
ductivity. Laboratory data showing the effect of closure stress on The initial closure stress in the Barnett example is 1,200 psi before
unpropped- and partially-propped-fracture conductivity for rock production (i.e., no drawdown), with maximum closures stress of
with relatively high Young’s modulus was published by Fredd 4,000 psi when the pressure in the network fractures reaches the
et al. (2001). Fredd’s laboratory data were extrapolated to lower FBHP of 1,000 psi. In reference to Fig. 10, the conductivity of the
Young’s modulus rock for unpropped fractures with shear offsets network fractures is approximately 0.55 md-ft when closure stress
to illustrate the impact of modulus on unpropped-fracture conduc- is 4,000 psi, almost a 20-fold decrease from the initial conductiv-
tivity (Cipolla et al. 2008a). These data and this approach were ity of 10 md-ft.
used to evaluate stress-sensitive network-fracture conductivity in Figs. 11 and 12 show the impact of stress-dependent fracture
the Barnett shale. Fig. 10 shows an estimation of the impact of conductivity on well performance (including desorbed gas). In
closure stress and Young’s modulus on unpropped-fracture con- these cases, the primary-fracture spacing is 600 ft and network-
ductivity, illustrating the dramatic decrease in fracture conductivity fracture spacing is varied from 50 to 200 ft, while matrix perme-
as closure stress increases and Young’s modulus decreases. In this ability is varied from 10−5 md to 10−4 md. The solid lines represent
example, a reference conductivity of 10 md-ft at 1,200-psi closure the cases without stress effect and with 2 md-ft constant fracture
stress was used for all cases. The reference conductivity of 10 md- conductivity, while the dashed lines represent the cases with stress
ft was found to be a reasonable estimate for the Barnett shale (E ≈ effect (assuming E = 5 10+6 psi) and 10-md-ft initial conductivity
5 10+6 psi) on the basis of comparisons to actual well performance, (Fig. 10). In this case, the impact of variable closure stress in the
but would actually be a history-matching parameter and could network fractures reduces the ultimate gas production by approxi-
vary considerably depending on stimulation designs, completion mately 8 to 10% and may alter the production profile.
strategies, and geologic variability. The same reference conductiv- The production profiles for different fracture spacings show that
ity was used for lower-modulus rock to allow easy comparison of stress-dependent fracture conductivity results in materially higher
the simulation results for lower-modulus rock and/or the impact of initial production rates when block sizes are small, while there
more-severely stress-dependent fracture conductivity. is very little impact on the initial production profile when block
Closure stress on the fracture is defined as the horizontal stress sizes are larger. In all cases, stress-dependent fracture conductiv-
perpendicular to the fracture (h) minus the pressure inside the ity reduces ultimate gas production, as drawdown in the fracture
fracture (pf). For simplicity, the two horizontal stresses are assumed network continues to increase throughout the well life and network
to be equal. In the Barnett shale, a horizontal stress of 5,000 psi fractures “close.” However, the impact of changing closure stress
on well production may not be critical in the Barnett shale owing
to the relatively high Young’s modulus. However, additional res-
ervoir-modeling work is required to better understand the impact
of stress-dependent fracture conductivity.
Comparing the production profile for Well B to the simulation
results in Figs. 11 and 12 shows that Well B follows the produc-
tion trend for the case with stress-dependent fracture conductivity,
10−4 md matrix permeability, and 600-ft network-fracture spacing
slightly better than the same case without stress-dependent fracture
conductivity (Fig. 11, lower set of curves). More-detailed history
matching should provide a better, but still not unique, match for
both wells.
Impact of Young’s Modulus and Severity of Stress-Dependent
Fracture Conductivity. As Young’s modulus decreases (i.e., softer
rock), the impact of increasing closure stress on unpropped- or par-
tially-propped-fracture conductivity will likely be more severe. An
approximation of the impact of Young’s modulus on unpropped-
fracture conductivity is shown in Fig. 10, illustrating the dramatic
decrease in unpropped-fracture conductivity that is likely for softer
rocks. A series of reservoir simulations was performed using the
Fig. 12—Impact of stress-dependent fracture conductivity for Barnett shale reservoir properties and the fracture-conductivity
the desorbed-gas case (k = 10−5 md). profiles in Fig. 10. Figs. 13 and 14 show the simulation results

August 2010 SPE Reservoir Evaluation & Engineering 643


Fig. 13—Impact of Young’s modulus on stress-dependent frac- Fig. 14—Impact of Young’s modulus on stress-dependent frac-
ture conductivity for the desorbed-gas case (k = 10−4 md). ture conductivity for the desorbed-gas case (k = 10−5 md).

for matrix permeability of 10−4 and 10−5 md, respectively. The stress-dependent network-fracture conductivity does not appear
simulations assume a primary- or main- fracture spacing of 600 ft to be significant, mostly affecting the mid- to late-life production
and a network-fracture spacing of 100 ft (k=10−4 md) and 50 ft behavior, and could reduce gas recovery by approximately 10%.
(k=10−5 md). The cumulative production for a constant and uni- However, if the relationship between closure stress and unpropped-
form 2-md-ft network-fracture conductivity and is shown for fracture conductivity is more severe than that shown in Fig. 10
comparison. The results show that more-severe stress-dependent (E = 5×106 psi curve), then ultimate gas recovery could be affected
fracture conductivity can reduce gas recovery significantly. When materially.
Young’s modulus is 4×106 psi, stress-dependent fracture conduc-
tivity reduces gas recovery by 11 to 14% in comparison with the Marcellus Shale: Impact of Stress-Dependent Network-Fracture
constant-conductivity case. For lower-modulus rock (2×106 psi), Permeability. The impact of gas desorption and stress-dependent
gas recovery decreases by 40 to 43% compared to the E=5×106 psi fracture conductivity was also evaluated for typical Marcellus shale
case. At lower modulus, the fracture conductivity is reduced sig- reservoir properties, although the Marcellus is more variable in
nificantly as drawdown in the fracture network increases with time, depth and thickness and much earlier in its development than the
resulting in poor drainage of the tight matrix rock and significantly core area of Barnett—making it more difficult to define average
lower ultimate gas recovery. However, initial well performance is or generic properties. In addition, actual production data were not
not markedly different for the various cases, and the severity of available to “calibrate” the Marcellus simulations. Table 2 lists the
stress-dependent fracture conductivity may not be evident during generic reservoir properties used for the Marcellus shale. The area
the first 1–2 years of production, requiring 3–5 years of produc- of the reservoir model is the same as that used in the Barnett shale,
tion history to better define the possible effects of stress-dependent 3,400×2,400 ft (≈190 acres), and a cased-and-cemented horizontal
network-fracture conductivity. completion with multiple fracture-treatment stages is assumed. The
The production profile for Well A follows the trend for a matrix SRV for the Marcellus simulations is 680×106 ft3 (100 ft of pay).
permeability of 10−5 md and a fracture-network spacing of 50 ft Fig. 15 shows a graph of gas content (scf/ton) vs. pressure
(lowest curve in Fig. 14) very closely, while also following the for the gas in the matrix porosity and the adsorbed gas used for
trend for a matrix permeability of 10−4 and network-fracture spac- the Marcellus shale (Faraj and Duggan 2008). The free gas in the
ing of 100 ft (lowest curve in Fig. 13). This illustrates some of the matrix porosity is shown as a nearly straight line (circles), while
issues with potentially non-unique solutions when history matching the adsorbed gas is represented by a curved line (squares). In this
shale-gas production profiles—because fracture conductivity, matrix case, 46.5% of the total gas is adsorbed, a percentage similar to
permeability, and network-fracture spacing are many times not well that in the Barnett shale. At higher pressures, more gas is stored
known. However, reservoir modeling can provide important insights in the matrix porosity and the gas porosity will contribute more
into production mechanisms and stimulation performance when to production than gas desorption because of the difference in the
values of matrix permeability and unpropped-fracture conductivity curve shapes (straight line vs. curved line).
can be bounded using laboratory data. It may be possible to estimate The Marcellus simulations include the effect of stress depend-
network-fracture spacing and network structure using MS-mapping ent network-fracture conductivity. Assuming uniform and constant
data, further constraining the history match.
Summary of Barnett Shale-Gas Desorption and Stress-Sensi-
tive-Permeability Simulations. The simulation study using typical TABLE 2—GENERIC MARCELLUS SHALE PROPERTIES
Barnett shale reservoir and rock properties indicated that gas
Depth (ft) 5,000
desorption and stress-dependent fracture conductivity probably
have a minor, but not necessarily insignificant, impact on well Reservoir pressure (psi) 2,750
performance. The impact of gas desorption is primarily in the Pore pressure gradient (psi/ft) 0.55
later life of the well when pressures in the tight matrix blocks may Closure pressure (psi) 3,500
become low enough to produce meaningful amounts of gas, result- Closure pressure gradient (psi/ft) 0.70
ing in an increase in 30-year gas recovery of 5–15%. However, as
Net pay (ft) 100
network-block size increases and fracture conductivity and matrix
permeability decrease, the ability to produce adsorbed gas becomes Porosity (% ) 8
increasingly difficult. While decreasing FBHP from 1,000 to 500 Matrix permeability (md) 1 E 4; 5 E 4
psi does not significantly increase the production of adsorbed Water saturation (%) 25
gas, it will improve 30-year gas recovery by approximately 10%
Reservoir temperature (°F) 105
(mostly because of the additional production of free gas). Because
of the relatively high modulus of the Barnett shale, the impact of FBHP (psi) 500

644 August 2010 SPE Reservoir Evaluation & Engineering


10
E=2E+6 psi

Conductivity (mD-ft)
1
2 mD-ft
0.02 mD-ft
0.1

0.01
Initial closure stress Closure stress at 500 psi FBHP

0.001
0 500 1000 1500 2000 2500 3000 3500 4000
Stress (psi)

Fig. 16—Effect of closure stress on unpropped-fracture con-


ductivity, Marcellus shale example.
Fig. 15—Example Marcellus shale-gas content, adsorbed and
free gas.
Thus, the impact of desorption decreases when the primary-frac-
ture spacing is larger or farther apart (i.e., fewer fracture-treatment
network-fracture conductivity may be optimistic for lower-modu- stages along the lateral). Free gas (porosity) is likely the dominant
lus shale, such as the Marcellus (E≈2×106 psi). Fig. 16 shows an contributor to production in the Marcellus shale.
estimate of the effect of closure stress on unpropped-fracture con- Fig. 19 shows pressure distribution after 1, 5, and 15 years for
ductivity for a Young’s modulus of 2 MMpsi. It should be empha- the two different matrix-permeability cases. Because of the sym-
sized that this is only an estimate based on previously published metry of the reservoir area investigated, only one-half of the total
work (Fredd et al. 2001; Cipolla et al. 2008a) and revised on the reservoir size is simulated. The primary fractures extend 1,000 ft
basis of Barnett shale history-matching results, which indicated from the wellbore (half-length). The fracture conductivity is 2 md-
that the previously published estimates may be optimistic. The ft for both the primary and network fractures (uniform conductiv-
initial network-fracture conductivity is 2 md-ft before production, ity). The pressure distribution includes the effect of desorption and
but declines to 0.02 md-ft when the pressure in the network frac- stress-dependent fracture conductivity. Dark red indicates initial
tures decreases to the FBHP of 500 psi. reservoir pressure, while dark blue indicates FBHP (500 psi). The
The Marcellus simulations assume that a higher-conductivity left side of the graph shows the pressure distribution for a matrix
primary fracture is present with an initial conductivity of 50 md-ft. permeability of 5×10−4 md, primary-fracture spacing of 300 ft, and
The conductivity of the primary fracture is also stress dependent, a network block size (Dx) of 300 ft. When matrix permeability is
decreasing to approximately 0.25 md-ft at 3,000-psi closure stress very low and network block size is large, pressures in the matrix
(when the pressure in the primary fracture is 500 psi). The impact blocks remain relatively high even after 15 years of production.
of a high-conductivity primary fracture will be discussed later in The right side of the graph shows the pressure distribution for a
this section. matrix permeability of 10−4 md, primary-fracture spacing of 100 ft,
Figs. 17 and 18 show the impact of gas desorption on cumula- and network-block size of 50 ft, showing that when network-block
tive gas production. In Fig. 17 (k = 10−4 md) and Fig. 18 (k = 5×10−5 size is small (i.e., more-complex fracture network), the tight matrix
md), the secondary or network-fracture spacing is fixed (100 ft) can be effectively drained after 15 years. In this comparison, matrix
and network fracture conductivity is 2 md-ft. The primary-fracture permeability is five times lower for the case with 50-ft network
spacing (Dm) is 100 and 300 ft. For the smaller primary-spacing blocks in comparison to the 300-ft case and gas recovery is still
case (100 ft apart), the desorbed gas contributes approximately significantly higher (as evidenced by the lower pressures). In many
10% of the total production after 30 years (for both matrix-perme- shale-gas reservoirs, gas recovery may be more dependent on the
ability cases). As was seen in the Barnett example, these graphs size of the network blocks (degree of fracture complexity) than
show that the desorbed gas contribution is relatively small within on the permeability of the matrix rock. However, because of the
the first 5 years of production. If the primary-fracture spacing is relatively high FBHP (500 psi) and the behavior of the Marcellus
larger (300 ft apart), the desorbed gas results in only 8% more desorption isotherm (Fig. 15), very little gas is desorbed from the
production after 30 years (for both matrix-permeability cases). shale even with efficient drainage of the tight matrix.

Fig. 17—Impact of desorbed-gas production in the Marcellus Fig. 18—Impact of desorbed-gas production in the Marcellus
shale (k = 10−4 md). shale (k = 5×10−4 md).

August 2010 SPE Reservoir Evaluation & Engineering 645


k = 5e-4 mD, primary spacing = 300 ft, Dx = 300 ft k = 1e-4 mD, primary spacing = 100 ft, Dx = 50 ft

1 year

5 years

15 years

Fig. 19—Pressure distribution after 1, 5, and 15 years for the generic Marcellus shale example (with desorbed gas).

Effect of Stress-Dependent Fracture Conductivity. The impact 10% additional gas production in 30 years, while having little effect
of stress-dependent conductivity is shown in Fig. 20 for a matrix on the initial production profile. Therefore, desorbed-gas produc-
permeability of 5×10−4 md, primary-fracture spacing of 100 ft, tion is probably a minor component in the economic development
and a network-fracture spacing of 100 ft. All simulations include of the Marcellus shale. In contrast to the Barnett shale, the lower
gas desorption. The cumulative production for a constant 2-md-ft Young’s modulus of the Marcellus could mean that stress-depen-
uniform-conductivity network is shown by the top curve, indicating dent network-fracture conductivity will play a significant role in
a 30-year gas recovery of 8,342 MMscf. The bottom curve (dashed well performance and could substantially reduce initial production
line) shows the production for a uniform-conductivity network rates and ultimate gas recovery. However, the detrimental effects
with stress-dependent fracture conductivity (initial conductivity of stress-dependent network-fracture conductivity may be offset
of 2 md-ft, declining to 0.02 md-ft when the pressure in a net- by stimulation designs that result in high-relative-conductivity
work fracture reaches 500 psi, Fig. 16). The 30-year gas recovery primary fractures.
is reduced significantly when stress-dependent network-fracture
conductivity is modeled, decreasing by almost 60%. Vertical-Well Case Study
The detrimental effects of stress-dependent network-fracture The example well is a vertical well located in the core area of the
permeability can be offset, to some degree, by designing stimulation Barnett shale; it was fracture treated and mapped in 2001 (May-
treatments that create a high-relative-conductivity primary fracture. erhofer et al. 2006). MS mapping clearly illustrates that very large
The presence of a high-relative-conductivity primary fracture fracture networks are generated, with the predominant hydraulic-
significantly reduces network-fracture conductivity requirements fracture azimuth in the northeasterly direction and a secondary
(Cipolla et al. 2008a, 2009a). The impact of a primary fracture
with an initial conductivity of 50 md-ft is also shown in Fig. 20,
middle curve with higher initial production and 7200×106 ft3 cumu-
lative gas production. It should be emphasized that the conductiv-
ity of the primary fracture is also stress dependent, declining to
0.25 md-ft when the pressure in the primary fracture reaches
500 psi. The cumulative gas production after 30 years is roughly
doubled for the case with a high relative conductivity primary
fracture in comparison with the 2-md-ft uniform-fracture-conduc-
tivity case (with stress-dependent network-fracture conductivity)—
essentially offsetting the detrimental affects of stress-dependent
network-fracture conductivity. It is interesting to note that the case
with a high-relative-conductivity primary fracture (Fig. 20, middle
curve) exhibits much higher initial gas-production rates compared
to the 2-md-ft constant- and uniform-conductivity case (Fig. 20,
top curve), yet 30-year cumulative gas production is approximately
15% lower. These simulations emphasize the importance of under-
standing the stress dependence of network-fracture conductivity in
fracture design and production forecasting.
Summary of Marcellus Shale Gas-Desorption and Stress-
Sensitive-Permeability Simulations. The impact of desorbed-gas Fig. 20—Impact of stress-dependent fracture conductivity, with
production in the generic Marcellus reservoir-simulation cases was and without a high-relative-conductivity primary fracture, Mar-
similar to that for the Barnett shale, possibly adding approximately cellus shale example.

646 August 2010 SPE Reservoir Evaluation & Engineering


TABLE 3—EXAMPLE-WELL RESERVOIR PARAMETERS

Parameters V a lu e

Depth (ft) 7,000


Net thickness (ft) 415
Porosity (% ) 6
Water saturation (%) 30
Initial pressure (psi) 3,800
South/North (ft)

Fracture-Network Model
Temperature (°F) 180
Gas gravity 0.6
MS mapping

gas rate was used as a constraint (Fig. 22). The total well history
includes 4 years of production. Fig. 23 shows the history match of
flowing tubing and casing pressures vs. time. The model predicts
higher wellhead flowing pressures in late time, which tracks the
surface annulus pressure (open-ended tubing) and represents liquid
loading. The loading condition was confirmed by higher flow rates
after a foam cleanout. Although this match is not entirely unique,
the effective total-fracture-network length (sum of all fracture seg-
West/East (ft) ments open to flow in the two principal frac directions) is on the
order of 6,000 to 7,800 ft, with an average conductivity of 4 md-ft
Fig. 21—Fracture network as determined from MS mapping (matrix permeability was fixed at 0.0001 md or 100 nanodarcies,
is approximated with numerical network model (Mayerhofer which was estimated from core tests). The effective network width
et al. 2006). was modeled to be 180 ft (90-ft fracture spacing in the northeast
direction and four orthogonal stringers) with a network extent of
approximately 2,000 ft in the southwest direction. While the total
component orthogonal to that. One of the remaining questions effective network length (sum of all open segments) must fall
concerns how much of the generated network is actually effective within the 6,000- to 7,800-ft range, the aspect ratio of effective
during production. Fig. 21 shows the MS-mapping results. The network extent to width could vary to some degree. An alternative
total extent of the fracture network is at least 2,700 ft, with apparent match, which would also be within the MS-mapped network area,
asymmetric growth to the southwest. It appears that the northeast was achieved with a wider 360-ft network (180 ft fracture spacing)
wing of the network structure did not develop as effectively, on the and shorter 1,300-ft extent in the southwest direction. Decreasing
basis of created MS events (Note: The observation well was in a the matrix permeability by an order of magnitude (to 10 nanodar-
position that should have detected growth on the northeast side). cies) would result in a similar pressure history match; however, the
The network width is fairly small in this case (approximately 350 effective fracture-network length would increase to approximately
ft), and network height shows adequate coverage of the Lower 16,500 ft and result in closer fracture spacing of 60 ft with 360-ft
Barnett interval. The network structure (bold lines) in Fig. 21 network width and 2,700-ft total extent. The exact position, aspect
represents the effective network modeled in the reservoir simulator ratio of effective network width to length, and fracture spacing
as an approximation of the MS results. This is the final network within the network can be uniquely determined only with the help
structure used to history match well production and is a result of of multiwell production data.
several iterations that explored alternative answers. In short, the Fig. 24 shows the pressure distribution after 1 year of produc-
conclusion is that once matrix permeability is fixed, the resulting tion. Red represents initial reservoir pressure, and dark blue is
total effective network length (sum of all open segments) and approximately the FBHP. Fig. 24 illustrates that nanodarcy matrix
conductivity are fairly unique. permeability will not drain very far beyond the created fracture
Table 3 shows the reservoir input parameters used in the simu- network. Additionally, low fracture conductivities play an impor-
lation. Three parameters were varied to obtain the history match: tant role because they create higher pressure drops in the network
fracture length, fracture density, and fracture conductivity. Actual resulting in less-efficient drainage farther from the well.
Wellhead Pressure (psi)

Time (days)

Fig. 22—History match of flowing tubing and casing pressures, Fig. 23—Gas-flow rate was used as a constraint for production
open-ended tubing (Mayerhofer et al. 2006). modeling (Mayerhofer et al. 2006).

August 2010 SPE Reservoir Evaluation & Engineering 647


Conductivity=
4 md⋅ft

Fig. 24—Pressure distribution after 1 year of production (May-


erhofer et al. 2006).

Fig. 25—Blasingame analysis, Barnett shale (xf = 7,000 ft).


Advanced Decline-Curve Analysis. The production data for this
well were evaluated in Cipolla et al. (2008b). The Blasingame
production type-curve analyses of the example well are shown in economics in many moderate-to-deep shale-gas plays may be
Figs. 25 and 26, illustrating the non-uniqueness that is possible in insignificant.
unconventional reservoirs when reservoir permeability is uncertain 3. The impact of stress-dependent network-fracture conductivity
and fracture growth is not measured. Figs. 25 and 26 show two in higher-Young’s-modulus shale reservoirs, such as the Barnett
matches to the production from the referenced Barnett shale well: one shale, may be small. In these hard-rock shale-gas reservoirs,
indicating an effective fracture half-length of 7,000 ft, matrix perme- reductions in network-fracture conductivity as closure stress
ability of 5×10−6 md, and a drainage area of 69 acres and a second increases could reduce ultimate gas recovery by 10%, but may
match indicating an effective fracture half-length of only 130 ft, matrix not have a material impact on initial well productivity.
permeability of 4×10−3 md, and a drainage area of 8.5 acres. 4. Production from shale-gas reservoirs with lower Young’s modulus,
Although constraining the matrix permeability would result in such as the Marcellus and Haynesville shales, could be affected
less uncertainty in the type-curve match, the examples illustrate significantly by stress-dependent network-fracture conductivity,
how the assumption of a single planar fracture and the inability resulting in significantly lower gas recovery than anticipated on the
to describe the network-fracture system limit the application of basis of initial well productivity. Although the impact of increasing
type-curve analyses in unconventional gas reservoirs. It can be very closure stress on network fractures can be severe for lower-modulus
important when evaluating production data to determine, even qual- rock, the effects may not be evident during the initial 1–2 years of
itatively, the effects of matrix permeability and primary-fracture production, complicating production forecasting.
conductivity (plus the impact of network size, fracture complexity, 5. Fracture-treatment designs that create a high-relative conduc-
and network-fracture conductivity) on well performance. tivity primary fracture may offset, at least to some degree, the
The Barnett shale is an example in which selecting the appro- detrimental effects of stress-dependent or low network-fracture
priate diagnostic technologies was the key to success. In this case, conductivity, resulting in significant improvements in produc-
fracture growth is too complex to be modeled using commercial tion rates and gas recovery.
fracture models; therefore, treatment designs are based on the inte- 6. Recovery of adsorbed gas may be improved if more-complex
gration of MS data and well-performance evaluations. To evaluate fracture networks can be created (i.e., smaller network blocks).
well performance properly requires numerical reservoir modeling 7. Numerical-reservoir-simulation history matching can be non-
that captures the details of the complex fracture network, which unique in complex shale-gas reservoirs. However, the modeling
limits the application of type-curve production-data-analysis tech- can be reasonably constrained when estimates of matrix perme-
niques. The combination of MS mapping and numerical reservoir ability are available from special core analyses and MS-map-
modeling resulted in significant insights into stimulation designs ping data are gathered to determine SRV and provide insights
and completions strategies, leading to the application of large- into fracture-network development. Combining these data with
volume high-rate water-frac treatments with low concentrations laboratory measurements of unpropped- and partially-propped-
of small proppants and horizontal wells to maximize the fracture- fracture conductivity can significantly improve the insights
network size and complexity. gained from reservoir simulation.

Conclusions
The reservoir-simulation work focused on evaluating the effects of
gas desorption and stress-dependent network permeability in hori-
zontal completions using reservoir properties typical of the Barnett
and Marcellus shales, but the results and modeling approach should
be applicable to many other shale-gas reservoirs. The following
conclusions can be drawn from this work.
1. Gas desorption may not be a significant component of produc-
tion in many moderate-to-deep shale-gas reservoirs, such as the
Barnett, Marcellus, and Haynesville shales. Although adsorbed
gas may constitute 40–50% of the total gas in place, the ability
to produce the adsorbed gas is limited because of the ultratight
matrix rock, relatively high FBHP, and the desorption profile,
which requires relatively low pressures to produce significant
amounts of adsorbed gas.
2. Desorbed gas may account for 5–15% of ultimate gas produc-
tion, but the desorbed gas is produced primarily during the later
life of the well, having little impact on the initial well produc-
tivity. Therefore, the impact of gas desorption on development Fig. 26—Blasingame analysis, Barnett shale (xf = 130 ft).

648 August 2010 SPE Reservoir Evaluation & Engineering


8. Production-data analyses, such as advanced type-curve solutions, Fisher, M.K., Wright, C.A., Davidson, B.M., Goodwin, A.K., Fielder,
cannot adequately describe the complex fluid flow in shale- E.O., Buckler, W.S., and Steinberger, N.P. 2002. Integrating Fracture
gas reservoirs. However, these techniques may have important Mapping Technologies to Optimize Stimulations in the Barnett Shale.
qualitative applications. Paper SPE 77411 presented at the SPE Annual Technical Conference
9. Discrete numerical modeling of the fracture network and tight and Exhibition, San Antonio, Texas, USA, 29 September–2 October.
matrix rock is necessary to accurately simulate production doi: 10.2118/77441-MS.
from many shale-gas reservoirs. Numerical-reservoir-simulation Frantz, J.H. Jr., Williamson, J.R., Sawyer, W.K., Johnston, D., Waters, G.,
approaches that use dual-porosity/dual-permeability approxima- Moore, L.P., Macdonald, R.J., Pearcy, M., Ganpule, S.V., and March,
tions may not adequately capture the very long transient behav- K.S. 2005. Evaluating Barnett Shale Production Performance Using an
ior in the ultratight shale matrix blocks. Integrated Approach. Paper SPE 96917 presented at the SPE Annual
Technical Conference and Exhibition, Dallas, 9–12 October. doi:
Nomenclature 10.2118/96917-MS.
Dm = primary-fracture spacing, L Fredd, C.N., McConnell, S.B., Boney, C.L., and England, K.W. 2001.
Experimental Study of Fracture Conductivity for Water-Fracturing and
Dx = network-fracture spacing, square blocks, L
Conventional Fracturing Applications. SPE J. 6 (3): 288–298. SPE-
E = Young’s modulus 74138-PA. doi: 10.2118/74138-PA.
Fc = fracture conductivity, L3 Kazemi, H. 1969. Pressure Transient Analysis of Naturally Fractured
FCD = dimensionless fracture conductivity kf /kxf Reservoir With Uniform Fracture Distribution. SPE J. 9 (4): 451–462;
h = net pay, L Trans., AIME, 246. SPE-2156-A. doi: 10.2118/2156-A.
k = matrix permeability, L2 Mayerhofer, M.J., Lolon, E.P., Youngblood, J.E., and Heinze, J.R. 2006.
kf = fracture permeability, L2 Integration of Microseismic Fracture Mapping Results With Numeri-
kwf = fracture conductivity, L3 cal Fracture Network Production Modeling in the Barnett Shale.
pf = pressure in fracture Paper SPE 102103 presented at the SPE Annual Technical Conference
xf = hydraulic-fracture wing or half-length, L and Exhibition, San Antonio, Texas, USA, 24–27 September. doi:
10.2118/102103-MS.
xn = hydraulic-fracture-network width (from MS event pattern), L
Mayerhofer, M.J., Lolon, E.P., Warpinski, N.R., Cipolla, C.L., Walser, D., and
h = horizontal stress Rightmire, C.M. 2010. What Is Stimulated Reservoir Volume? SPE Prod
& Oper 25 (1): 89–98. SPE-119890-PA. doi: 10.2118/119890-PA.
Acknowledgments Medeiros, F., Ozkan, E., and Kazemi, H. 2008. Productivity and Drainage
The authors would like to thank Carbo Ceramics and StrataGen Area of Fracture Horizontal Wells in Tight Gas Reservoirs. SPE Res Eval
Engineering for supporting the publication of this work and Com- & Eng 11 (5): 902–911. SPE-108110-PA. doi: 10.2118/108110-PA.
puter Modelling Group for supporting the shale-gas reservoir-mod- Montgomery, S.L. 2004. Barnett Shale: A New Gas Play in the Fort Worth
eling studies. The authors would also like to acknowledge Michael Basin. IHS Energy Petroleum Frontiers 20 (1).
Mayerhofer for his many contributions to shale-gas reservoir Warpinski, N.R., Mayerhofer, M.J., Vincent, M.C., Cipolla, C.L., and Lolon,
modeling that were instrumental in this work. E.P. 2008. Stimulating Unconventional Reservoirs: Maximizing Network
Growth While Optimizing Fracture Conductivity. Paper SPE 114173
References presented at the SPE Unconventional Reservoirs Conference, Keystone,
Arthur, J.D., Bohm, B., and Layne, M. 2008. Hydraulic Fracturing Con- Colorado, USA, 10–12 February. doi: 10.2118/114173-MS.
siderations for Natural Gas Wells of the Marcellus Shale. Presented at Warren, J.E. and Root, P.J. 1963. The Behavior of Naturally Fractured
The Ground Water Protection Council 2008 Annual Forum, Cincinnati, Reservoirs. SPE J. 3 (3): 245–255; Trans., AIME, 228. SPE-426-PA.
Ohio, USA, 21–24 September. doi: 10.2118/426-PA.
Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J., Lolon, E.P., and Vincent,
M.C. 2008a. The Relationship Between Fracture Complexity, Reser- Appendix A—Reservoir Simulation, Shale-Gas
voir Properties, and Fracture Treatment Design. Paper SPE 115769 Gridding, and Modeling Issues
presented at the SPE Annual Technical Conference and Exhibition, Numerical modeling of gas production from hydraulically frac-
Denver, 21–24 September. doi: 10.2118/115769-MS. tured shale-gas wells has been studied to determine how best to
Cipolla, C.L., Lolon, E.P., and Mayerhofer, M.J. 2008b. Resolving Created, represent network and/or natural fractures (both inside and outside
Propped, and Effective Hydraulic Fracture Length. Paper IPTC 12147 the stimulated reservoir volumes) accurately and efficiently.
presented at the International Petroleum Technology Conference, Kuala
Lumpur, 3–5 December. doi: 10.2523/12147-MS. Modeling Single-Plane Propped Fractures in Single-Porosity
Cipolla, C.L., Lolon, E.P., Mayerhofer, M.J., and Warpinski, N.R. 2009a. The Reservoirs. This work extends previous work for modeling single-
Effect of Proppant Distribution and Un-Propped Fracture Conductivity on plane propped fractures in single-porosity reservoirs (e.g., tight
Well Performance in Unconventional Gas Reservoirs. Paper SPE 119368 gas sands) in which it was determined that transient, non-Darcy
presented at the SPE Hydraulic Fracturing Technology Conference, The flow could be accurately (in comparison to very finely gridded,
Woodlands, Texas, USA, 19–21 January. doi: 10.2118/119368-MS. single-porosity reference models and analytical solutions) and
Cipolla, C.L., Lolon, E.P., and Dzubin, B. 2009b. Evaluating Stimulation efficiently (run times of a couple of minutes vs. hours) modeled
Effectiveness in Unconventional Gas Reservoirs. Paper SPE 124843 with a finite-difference simulator using a locally refined grid: (a)
presented at the SPE Annual Technical Conference and Exhibition, whose innermost row of cells representing the propped fracture
New Orleans, 4–7 October. doi: 10.2118/124843-MS. can be orders of magnitude larger in width than the actual propped
Cipolla, C.L., Lolon, E.P., Mayerhofer, M.J., and Warpinski, N.R. 2009c. fracture width (i.e., ≈2 ft) as long as both the fracture permeability
Fracture Design Considerations in Horizontal Wells Drilled in Uncon- and the non-Darcy-flow coefficient are scaled appropriately, and
ventional Gas Reservoirs. Paper SPE 119366 presented at the SPE (b) whose spacing perpendicular to the length of the propped
Hydraulic Fracturing Technology Conference, The Woodlands, Texas, fracture is logarithmic. A graph of the grid design is shown in
USA, 19–21 January. doi: 10.2118/119366-MS. Fig. A-1. A comparison of FBHP behavior with this grid design
Faraj, B. and Duggan, J. 2008. Marcellus Shale Gas Potential in the South- (Fig. A-1) vs. a very finely gridded single-porosity reference
ern Tier of New York. Presented at the Annual Canadian Society for model and Gringarten’s infinite-conductivity type-curve solutions
Unconventional Gas (CSUG) Conference, Calgary, 19–21 November. is presented in Fig. A-2.
Fisher, M.K., Heinze, J.R., Harris, C.D., Davidson, B.M., Wright, C.A.,
and Dunn, K.P. 2004. Optimizing Horizontal Completion Techniques Modeling Fracture Networks. This section presents an extension
in the Barnett Shale Using Microseismic Fracture Mapping. Paper SPE of previous work on single-plane propped vertical fractures in
90051 presented at the SPE Annual Technical Conference and Exhibi- vertical and horizontal wells to the modeling of networks of frac-
tion, Houston, 26–29 September. doi: 10.2118/90051-MS. tures in shale-gas wells that have been completed with multistage

August 2010 SPE Reservoir Evaluation & Engineering 649


Detailed Gridding of Hydraulic Fracture (xf = 500 ft)

Hydraulic Fracture
Fig. A-1—Grid design for single-plane vertical-propped-fracture simulation.

fracture treatments. Several options were evaluated to determine The MINC approach for modeling the fracture network was
which “grid design” and “porosity model” approach resulted also evaluated but proved inadequate at early time, as can be seen
in the best agreement with very finely gridded, single-porosity in Fig. A-5, which compares logarithmically and equally spaced
reference models, including the use of the multiple interacting gridding for both MINC (with 4 levels of refinement) and the new
continuum (MINC) and the dual-permeability “dual-continuum” simplified 9×9 LGR grid with 2-ft-wide grids representing the
models, while minimizing model run time. fractures. The MINC approach provided worse results when flow
Our work has determined that the best method for both accu- was allowed from the shale surrounding the stimulated volumes
rately and efficiently modeling transient, non-Darcy gas produc- (not shown).
tion from hydraulically fractured horizontal shale-gas wells is
to (a) use the dual-permeability method to represent all network Modeling Variations in Fracture Intensity in the Stimulated
fractures in both the unstimulated and stimulated volumes and Volume. We extended the new DK-LS-LGR approach to account
to (b) locally refine the grid in the stimulated volumes using the for variable fracture intensity in the stimulated volume by (a)
same logarithmically spaced grid design as was developed for the using a base gridblock size throughout the model whose length is
single-plane vertical-fracture work mentioned previously. This new set equal to the largest fracture spacing observed in a given situ-
approach is henceforth referred to in this paper as the “DK-LS- ation, (b) applying logarithmic-spaced LGR within the simulated
LGR method” (i.e., for dual permeability, logarithmically spaced, volumes, and (c) varying the dual-permeability parameters within
locally refined grid) for modeling hydraulically fractured shale-gas the stimulated volumes to account for variable fracture intensity
wells/reservoirs. (i.e., spacing) and effective permeability. Fig. A-6a depicts the
Fig. A-3 compares the grid designs within the stimulated vol- stimulated-volume region surrounding a horizontal shale-gas well
umes for both the very-fine-grid reference model (a), which uses as is typically interpreted from MS data gathered during hydraulic-
43×43 LGR with 0.001-ft-wide grids representing each fracture fracturing treatments. The base grid length in this example is 200 ft
and the simplified model (b), which uses 9×9 LGR with 2-ft-wide on a side and represents the coarsest fracture spacing observed in
grids representing each fracture. Both the reference grid and the this particular shale formation. Fig. A-6b shows how we apply log-
simplified grid use logarithmic spacing to capture transient flow arithmically spaced, local grid refinement to the stimulated-volume
behavior to and within the fractures. The gas-production rates from region, and also how we modify effective fracture permeability (the
these two grids are shown in Fig. A-4 and confirm that the simpli- green areas) and the fracture spacing (not shown) to represent vari-
fied approach gives an excellent match to the reference model. able-intensity fracturing within the stimulated volumes. Fig. A-7

100.00
xe/xf=3

xe/xf=7

xe/xf=15
10.00
Infinite

IMEX Infinite
PD

1.00 IMEX xe/xf-15

IMEX xe/xf-7

IMEX xe/xf=3

0.10

0.01
0.0001 0.001 0.01 0.1 1 10 100 1000 10000
tDxf
(a) (b)

Fig. A-2—FBHP comparison between Fig. A-1 grid design and (a) a very-finely gridded single-porosity reference model and (b)
Gringarten’s infinite-conductivity type-curve analytical solution.

650 August 2010 SPE Reservoir Evaluation & Engineering


Reference Grid Simplified Grid
4.0 md-ft fractured shale LGR model 4.0 md-ft fractured shale LGR model
Permeability J (md) 2000-01-01 K layer: 1 Effective Permeability J (md) 2010-01-01 K layer: 1
10,099.80 10,099.90 10,100.00 10,100.10
9,800 9,900 10,000 10,100 10,200 10,300 10,400

-9,900
-10,099.90

-10,099.90

-10,000

-10,000
Network Fractures
Network Fractures
-10,100.00

-10,100.00
Producer1

-10,100

-10,100
Producer1

-10,200
-10,200
-10,100.10

-10,100.10

-10,300
-10,300
0.00 75.00 150.00 feet
0.00 25.00 50.00 meter

10,099.80 10,099.90 10,100.00 10,100.10 9,800 9,900 10,000 10,100 10,200 10,300 10,400

(a) ~0.45 ft (b) ~700 ft

Fig. A-3—Grid designs for (a) reference grid (43×43 LGR with 0.001-ft-wide grid for each fracture) and (b) simplified grid (9×9
LGR with 2-ft-wide grid for each fracture) representing network of propped fractures in stimulated volume of a hydraulically
fractured shale-gas well.

compares the production from a very-fine-grid single-porosity • Fractures of greater or variable intensity inside the stimulated
reference model and the DK-LS-LGR model (Fig. A-6) for shale volumes are modeled using scaled dual permeability properties
matrix permeability of (a) 0.0001 md and (b) 0.01 md, showing that (matrix and fracture permeability and fracture spacing).
the simplified gridding approach provides acceptable accuracy. • Outside the stimulated volume the grid is not refined.
The new approach to modeling hydraulically fractured shale- The advantages of the approach described in this paper
gas wells can be summarized as follows: for modeling hydraulically fractured shale-gas wells are as
• A dual-permeability approach has been validated for mod- follows:
eling the matrix and natural fractures both inside and outside of the • Dual-permeability behavior is accounted for both inside and
stimulated volumes in hydraulically fractured shale-gas wells. outside the stimulated volumes (important in some shale plays in
• The base fracture network in the stimulated volumes is mod- which unstimulated fracture permeability may be significant).
eled with 9×9 locally refined, logarithmically spaced grids in the • Variations in fracture intensity can be accommodated without
stimulated volumes. having to vary the base gridblock size.

Noflow Outside Stimulated Region


Pseudoization Test for Fine Grid
2.87e+6

2.37e+6 Fracture Pseudoized using 2.000 ft thick blocks


Fracture Modelled using 0.001 ft thick blocks
Gas Rate SC (ft3/day)

1.87e+6

1.37e+6

8.71e+5

3.71e+5
0 10 20 30 40 50 60
Time (day)

Fig. A-4—Gas-production-rate comparison between Fig. A-3 grid designs for the case of 0.0001-md matrix permeability.

August 2010 SPE Reservoir Evaluation & Engineering 651


Noflow Outside Stimulated Region Noflow Outside Stimulated Region
First 30 Days Equally and Log Spaced MINC vs. Equally and Log Spaced LGR
Equally and Log Spaced MINC vs. Equally and Log Spaced LGR 3.00e+6
3.00e+6

2.50e+6
2.50e+6

Gas Rate SC (ft3/day)


Gas Rate SC (ft3/day)

2.00e+6
2.00e+6 9x9 LGR 2 ft Bf Other Cells Equally Spaced
9x9 LGR 2 ft Bf Log Scaled
Minc 4 Equally Spaced
1.50e+6 1.50e+6 Minc 4 Log Scaled

1.00e+6 1.00e+6
9x9 LGR 2 ft Bf Other Cells Equally Spaced
9x9 LGR 2 ft Bf Log Scaled
5.00e+5 Minc 4 Equally Spaced 5.00e+5
Minc 4 Log Scaled

0.00e+0 0.00e+0
0 10 20 30 0 1,000 2,000 3,000 4,000
Time (day)
Time (day)
(a) (b)

Fig. A-5—Comparison of MINC dual-porosity approach to 9×9 LGR with 2-ft-wide grid representing each fracture.

Fig. A-6—(a) Typical stimulated volume resulting from a hydraulic-fracture treatment in a shale-gas well, and (b) the DK-LS-LGR
grid design created to model gas-production rate accurately in a numerical reservoir simulator.

New Dual Perm Model (0.0001 mD Shale) New Dual Perm Model (0.01 mD Shale)
Using Variable Fracture Spacing to Represent Using Variable Fracture Spacing to Represent
Different Fracture Intensity Different Fracture Intensity
Representing 66 ft Fracture Spacing in 200 ft DK blocks Representing 66 ft Fracture Spacing in 200 ft DK blocks
4.00e+6 4.00e+6
Single Porosity Run (Explicit Modelling Single Porosity Run (Explicit Modelling
or Variable Fracture Spacing or Variable Fracture Spacing
DK Run (Using Fracture Spacing DK Run (Using Fracture Spacing
3.00e+6 Scaling or 200 ft Blocks) 3.00e+6
Gas Rate SC (ft3/day)
Gas Rate SC (ft3/day)

Scaling or 200 ft Blocks)

2.00e+6 2.00e+6

1.00e+6 1.00e+6

0.00e+0 0.00e+0
2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010
Time (day) Time (day)
(a) (b)

Fig. A-7—Gas-production-rate comparison between Fig. A-6 grid design and a very finely gridded single-porosity reference model
for matrix-permeability cases of 0.0001 and 0.01 md.

652 August 2010 SPE Reservoir Evaluation & Engineering


and chemistry from the U. of Nevada-Las Vegas and a Masters
SI Metric Conversion Factors degree in petroleum engineering from the U. of Houston.
Elyezer Lolon is a reservoir engineer consultant for the Reservoir
acre × 4.046 873 E+03 = m2 Technology Group at Anadarko Petroleum Corporation, focus-
bbl × 1.589 873 E–01 = m3 ing on reservoir modeling studies for shale plays. Previously, he
ft × 3.048* E–01 = m worked for StrataGen Engineering and Pinnacle Technologies
ft2 × 9.290 304* E–02 = m2 in Houston and Chevron Pacific Indonesia, where he designed
ft3 × 2.831 685 E–02 = m3 and evaluated hydraulic fracturing treatments, analyzed diag-
nostic fracture injection tests, performed pressure transient and
gal × 3.785 412 E–03 = m3 production data analyses, and conducted integrated reservoir
lbm × 4.535 924 E–01 = kg studies. Lolon holds a BS degree in petroleum engineering from
psi × 6.894 757 E+00 = kPa Bandung Institute of Technology in Indonesia and MS and Ph.D.
degrees in petroleum engineering from Texas A&M U. J.C. Erdle
*Conversion factor is exact. is Computer Modelling Group’s vice president for software sales
and support for the USA and Latin America. He has 35 years
of industry experience, primarily in reservoir and production
Craig L. Cipolla is an engineering advisor for Schlumberger in engineering-related positions within the services and software
Houston, currently focusing on the application of MS-fracture segments of the E&P industry. Erdle holds BS (1971) and PhD
mapping to improve stimulation designs and field development (1974) degrees from Penn State in petroleum engineering. Barry
strategies. Before joining Schlumberger in 2009, his most recent Rubin is a senior staff engineer at Computer Modelling Group in
positions were VP of stimulation technology for Carbo Ceramics Calgary. He has worked in the area of simulator development
(2008–2009) and VP of engineering for Pinnacle Technologies at the Computer Modelling Group in Calgary for 20 years and
(1996–2008). Cipolla’s 29 years of worldwide experience at BP Canada, BP Exploration, and BP Research International for
includes the application of MS and tilt-meter fracture mapping a total of 12 years. Rubin’s research interests include the mod-
technologies, the design and evaluation of hydraulic fractur- eling of naturally fractured reservoirs and, more recently, the
ing treatments, tight gas reservoir engineering, integrated field modeling of non-Darcy gas flow in shales. He holds a Masters
studies, training, and supervising stimulation treatments. He was degree in chemical engineering from the U. of Calgary and
an SPE Distinguished Lecturer on hydraulic fracturing in 2005– a Bachelors degree in chemical engineering from McGill U. in
2006. Cipolla holds undergraduate degrees in engineering Montreal, Canada.

August 2010 SPE Reservoir Evaluation & Engineering 653


Comparison of Fractured-Horizontal-Well
Performance in Tight Sand and Shale
Reservoirs
E. Ozkan, SPE, and M. Brown, SPE, Colorado School of Mines; R. Raghavan, SPE, ConocoPhillips (Retired);
and H. Kazemi, SPE, Colorado School of Mines

Summary As shown in Fig. 2, the Raghavan et al. (1997) correlation


This paper presents a discussion of fractured-horizontal-well per- includes multiple curves for different numbers of fractures along
formance in millidarcy permeability (conventional) and micro- to the well. The top curve in Fig. 2 represents the correlation for a
nanodarcy permeability (unconventional) reservoirs. It provides single fracture (Prats 1961; Cinco-Ley and Samaniego-V. 1981).
interpretations of the reasons to fracture horizontal wells in both Fig. 2 may be used in two ways. It can be used to determine the
types of formations. The objective of the paper is to highlight the effective (total) system conductivity (CFD,eff) for given individual
special productivity features of unconventional shale reservoirs. fracture properties (conductivity CFD and half-length xF), distance
By using a trilinear-flow model, it is shown that the drainage between the outermost fractures (D), and the number of fractures
volume of a multiple-fractured horizontal well in a shale reservoir along the well (nF). Then, using the effective system conductivity
is limited to the inner reservoir between the fractures. Unlike con- (CFD,eff) and half-length (xF,eff=D/2), effective wellbore radius (rw,eff)
ventional reservoirs, high reservoir permeability and high hydrau- of the (total) system is obtained from the single-fracture correla-
lic-fracture conductivity may not warrant favorable productivity tion (the top curve in Fig. 2). Alternatively, we may use Fig. 2 to
in shale reservoirs. An efficient way to improve the productivity estimate the conductivity of the individual fractures (CFD) required
of ultratight shale formations is to increase the density of natural for a desired effective wellbore radius (rw,eff) for a given D and n.
fractures. High natural-fracture conductivities may not necessar- As noted by Chen and Raghavan (1997), the total effective
ily contribute to productivity either. Decreasing hydraulic-fracture wellbore radius can be used in the following well-known deliver-
spacing increases the productivity of the well, but the incremental ability equation:
production gain for each additional hydraulic fracture decreases.
The trilinear-flow model presented in this work and the informa- qsc =
7.08 × 10 −3 kh (
p − pwf ) . . . . . . . . . . . . . . . . . . . (1)
tion derived from it should help the design and performance predic- B ⎛1 4A ⎞
tion of multiple-fractured horizontal wells in shale reservoirs. ln
⎜ 2 e C r 2 ⎟
⎝ A w ,eff ⎠

Introduction Then, using Fig. 2 and Eq. 1, it is possible to design a multiple-


The objective of hydraulic fracturing in conventional tight reser- fractured horizontal well to accomplish the desired productivity.
voirs (as in tight sands with permeabilities in the milli- to micro- For example, for given hydraulic-fracture properties, it is possible
darcy range) has been to create a high-conductivity flow path to to optimize the number of fractures along a given horizontal-well
improve flow convergence in the reservoir. Accordingly, a practical length. Similarly, the optimum number of fractures, the length of the
interpretation of the objective of fracturing a horizontal well is to horizontal well, and the properties of the hydraulic fractures required
create a system whose long-term performance is identical to that to drain a given reservoir area of A (with the shape determined by
of a single effective (total) fracture of length equal to the spacing the shape factor CA) can be found from Fig. 2 and Eq. 1.
between the outermost fractures (Raghavan et al. 1997; Chen and It must be emphasized, however, that this approach is appropri-
Raghavan 1997). Fig. 1 provides a sketch of this interpretation. ate for the long-term performances of multiple-fractured horizontal
With this interpretation, performances of fractured horizontal wells wells dominated by flow beyond the tips of the fractures and the
can be correlated in terms of an effective fracture conductivity horizontal well. In other words, pseudoradial-flow convergence
(CF,eff) and effective fracture half-length (xF,eff). The conductivity of must have developed around the multiple-fractured horizontal well
this effective fracture depends on the permeability of the reservoir for a constant effective wellbore radius to be defined. Pseudoradial
and the number of, distance between, and conductivities of the flow develops around a single infinite-conductivity vertical fracture
individual hydraulic fractures. for times (Gringarten et al. 1974)
Raghavan et al. (1997) have presented practical results to evalu-
ate the long-term performances of multiple-fractured horizontal
1.14 × 10 4 ct  x F2
wells in low-permeability, conventional formations. Assuming t≥ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
equally spaced fractures of identical properties, they correlated k
the long-term performances (after the onset of pseudoradial flow)
of multiple-fractured horizontal wells in terms of the effective For long fractures (large xF) and tight formations (small k), time
wellbore radius of individual fractures (normalized by the dis- for the start of pseudoradial flow may be very long. For a mul-
tance between the outermost fractures, rwj/D) vs. effective (total) tiple-fractured horizontal well, because xF in Eq. 2 corresponds to
system conductivity (CFD,eff). The Raghavan et al. (1997) correla- xF,eff=D/2, the start of pseudoradial flow may be delayed even for
tion, shown in Fig. 2, is similar to the correlation presented by moderate-to-large permeabilities. Therefore, most of the produc-
Prats (1961) and Cinco-Ley and Samaniego-V. (1981) to obtain tive life of a fractured horizontal well is dominated by transient
the effective wellbore radius of a single fracture as a function of production from the stimulated volume between the hydraulic
fracture conductivity and half-length. fractures (Raghavan et al. 1997). In fact, pseudoradial flow may
not be applicable in unconventional tight reservoirs with micro- to
nanodarcy permeability (as in shale) in which the contribution of
Copyright © 2011 Society of Petroleum Engineers
the reservoir beyond the stimulated volume is usually negligible
This paper (SPE 121290) was accepted for presentation at the Western North American (Fisher et al. 2004; Mayerhofer et al. 2005; Maxwell et al. 2009;
Regional Meeting, Anaheim, California, USA, 26–30 May 2010, and revised for publication.
Original manuscript received for review 26 January 2009. Revised manuscript received for
Medeiros et al. 2008). Under these conditions, flow convergence
review 11 October 2010. Paper peer approved 11 October 2010. is mainly in the linear direction perpendicular to the surfaces

248 April 2011 SPE Reservoir Evaluation & Engineering


r w /xf vsCCFFD
D

r wj /D
Fig. 1—Multiple-fractured horizontal well in a conventional tight
formation and the effective (total) fracture concept. CFD,
FD,eff
,eff

of the hydraulic fractures and the long-term performance of a Fig. 2—Effective wellbore radius vs. effective system conduc-
fractured-horizontal-well/reservoir system can be represented by tivity for a multiple-fractured horizontal well in a conventional
that of a single hydraulic fracture. The length of the equivalent tight reservoir [after Raghavan et al. (1997)].
hydraulic fracture is equal to the aggregate length of the hydraulic
fractures, and the conductivity of the equivalent fracture is equal formation is dominated by linear-flow regimes. Our interpretation
to the average of the conductivities of the individual fractures, as is similar to the ideas used for fractured vertical wells in commin-
shown in Fig. 3. The drainage volume around the equivalent frac- gled reservoirs by Camacho-V. (1984), Bennett et al. (1985), and
ture is a parallelepiped with the length of the equivalent fracture Camacho-V. et al. (1987). As shown by the schematic in Fig. 4,
(2xF,eff=2nFxF) and the width equal to the average hydraulic-fracture the trilinear-flow model couples linear flows in three contiguous
spacing (dF). flow regions: the outer reservoir (denoted by the subscript O), the
In the following sections, we will introduce a trilinear-flow inner reservoir between fractures (denoted by the subscript I), and
model based on the premise that the drainage volume is limited the hydraulic fracture (denoted by the subscript F). In this work,
to the reservoir volume between hydraulic fractures. In extremely we assume uniform distribution of identical hydraulic fractures
tight formations, such as shale, an invasive (natural or induced) along the length of the horizontal well. (If fractures are not equally
fracture network is required for the efficient drainage of the matrix. spaced, then the trilinear model is applicable before the interaction
This introduces additional flow characteristics without changing of the hydraulic fractures.) Although it is possible to incorporate
the predominantly linear-convergence of flow from reservoir to the individual hydraulic-fracture properties in the model by an
hydraulic fractures. Because the linear flow performances of approach similar to that of Raghavan et al. (1997), the common
multiple-fractured horizontal wells in conventional and unconven- field practice of creating equally spaced hydraulic fractures of
tional tight formations can be evaluated along the same lines, we similar properties justifies our assumption. Furthermore, unless
will discuss them together. First, we will introduce the analytical there is significant contrast in individual-fracture properties, our
trilinear-flow model used in this work to simulate the performances model should closely approximate the performance of the system
of multiple-fractured horizontal wells. Then we will use it for the when average fracture properties are used.
numerical computations and demonstration examples. To simulate the performances of multiple-fractured horizontal
wells in shale, we allow the inner reservoir to be naturally frac-
Trilinear-Flow Model tured. We use the dual-porosity idealization to represent the natu-
The basis of our trilinear-flow model is the premise that the rally fractured medium. In this paper, we consider transient fluid
productive life of a multiple-fractured horizontal well in a tight transfer from the matrix (m) to natural fractures (f) as idealized by
Kazemi (1969), de Swaan O (1976), and Serra et al. (1983). The
model, however, is general and can be used with the pseudosteady-
fluid-transfer formulation of Warren and Root (1963). Our choice
of the transient dual-porosity model in this paper is based on the
expectation of elongated transient flow periods in tight matrices.
This expectation is supported by the quarter-slope behavior of the
log-log pressure response of pressure-transient tests observed in
the field, which is a result of the intermediate-time behavior of the
transient-fluid-transfer model. We discuss the flow-regime charac-
teristics of transient and pseudosteady dual-porosity models and
their implications for pressure-transient analysis in a companion
paper (Brown et al. 2009). We note only that the pseudosteady-
fluid-transfer model may be applicable in formations with matrix
permeabilities much higher than those observed in common appli-
cations of multiple-fractured horizontal wells.
The derivation and verification of the trilinear model are given
by Brown (2009) and Brown et al. (2009). Final analytical expres-
sions are presented in Appendix A. For our discussions, here we
note some essential definitions. We let qF,sc denote the flow rate
for each hydraulic fracture such that qF,sc = qsc/nF, where qsc is the
total flow rate from the horizontal well and nF is the number of
Fig. 3—Effective-fracture concept for a multiple-fractured hori- hydraulic fractures along the well length. For a naturally fractured
zontal well in an unconventional tight (shale) reservoir. inner zone, we use the intrinsic properties of the matrix and the

April 2011 SPE Reservoir Evaluation & Engineering 249


Line of symmetry
No-Flow Line

y e =d F /2 y e =d F /2
No-Flow Boundary

OUTER RESERVOIR OUTER RESERVOIR


HYDRAULIC kOO, φO,, cctO HYDRAULIC
tO kOO, φO,, cctO
tO
FRACTURE FRACTURE
k F, w FF, φF, c tF k F, w F φ
F, F, ctF
tF

xe xe
INNER RESERVOIR INNER RESERVOIR
NATURALLY NATURALLY
xF FRACTURED xF FRACTURED
k f, φf, c tf , k f, φf, c tf ,
k m, φm , c tm km, φm,, cctm tm

Line of symmetry
No-Flow Line
HORIZONTAL WELL HORIZONTAL WELL

Conceptual Representation of the Trilinear Model Symmetry Element Considered in the Mathematical Model
1/nF of the Well’s Drainage Area 1/4 of the Fracture’s Drainage Area
Producing q/nF stb/d (oil) or Mscf/d (gas) Producing qF /4 =q/(4nF) stb/d (oil) or Mscf/d (gas)

Fig. 4—Schematic of the trilinear-flow model used for the analytical solution of multiple-fractured-horizontal-well performance.

fractures. Specifically, kIh = kf hf , where kf is the intrinsic perme- convergence in fractures is radial around the wellbore intersection.
ability of the natural fractures, and hft is the total thickness of the However, unless the fracture length (2xF) is comparable to fracture
natural fractures intersecting the height of the hydraulic fracture height, flow within the fracture is predominantly linear, except
hF (in this work, hF is equal to the pay thickness h). If nf is the with the radial-flow convergence around the wellbore that causes
count of natural fractures intersecting the height of the hydraulic a choking effect (Soliman et al. 1990; Larsen and Hegre 1994;
fracture and hf is the aperture of the natural fractures, we have hft = Al-Kobaisi et al. 2006). We account for the effect of flow choking
nf hf . We also introduce the density of natural fractures, f = nf /h, around the intersection of the horizontal well and hydraulic frac-
in units of the number of natural fractures per foot of pay zone ture by incorporating a choking-skin (sc) expression (see Appendix
(f/ft). An estimate of f can be obtained from cores, logs, rate of A, Eq. A-23) given by Mukherjee and Economides (1991). The
penetration, and other sources. choking-skin concept should be valid after the end of radial flow
For the clarity of the discussions presented later, it is useful in the fracture. Early-time radial flow within hydraulic fractures
to comment on the no-flow boundaries and the lines of symmetry does not last long and is masked by the effect of wellbore stor-
considered in the trilinear model. As shown in Fig. 4, for the age. For the purposes of this work, we do not consider the effect
assumptions of the trilinear model, the lines bisecting the width of wellbore storage in our discussions. However, wellbore-storage
of each fracture and the lines bisecting the distance between two effect can be incorporated easily into the trilinear model by using
identical fractures (dF) are no-flow lines of symmetry. Similarly, Eq. A-24 in Appendix A. For the discussion of wellbore-storage
the axis of the horizontal well forms a no-flow line of symmetry. effect on trilinear-model responses, we refer the reader to Brown
Therefore, it is sufficient to model a symmetry element, which is et al. (2009).
one-quarter of a fracture’s drainage area with a production rate of
qF,sc/4=qsc/(4nF). In this model, the effect of the no-flow bound- Effect of the Outer Reservoir
ary at ye = dF/2 corresponds to the effect of interference between In Figs. 5 and 6, we present some results to support the premise
hydraulic fractures. and field observations that the drainage of fractured horizontal
We also consider another no-flow boundary at a distance xe ≥ wells is confined to the inner reservoir between the hydraulic frac-
xF from the axis of the horizontal well. The boundary at xe may tures (also referred to as the stimulated volume). Before discussing
correspond to a no-flow boundary between two wells or to a physi- the results, we must comment on the value of the matrix perme-
cal barrier to flow. Assuming xe = xF limits the drainage area to the ability (km=1,000 md) used in Figs. 5 and 6. When we used matrix
stimulated inner reservoir between hydraulic fractures. If xe > xF permeabilities in the millidarcy range in our simulations, we did
and the outer reservoir has considerably lower permeability com- not observe any influence of the outer reservoir because the outer
pared with the inner reservoir, then the flow across the boundary reservoir is made of the same matrix (without natural fractures).
between the inner and the outer reservoirs is negligible initially; A tighter matrix causes faster pressure drop in the inner zone,
that is, x = xF is a no-flow boundary at early times. The outer res- but the support of the outer zone is also less efficient and takes a
ervoir may contribute to flow when pressure drops significantly to longer period to be felt. Therefore, the cases presented in Figs. 5
start depletion in the inner reservoir. The magnitude and the time and 6 may not correspond to tight, unconventional reservoirs but
of contribution of the outer reservoir depend on the permeability they are useful to make the pertinent observations on the general
and volume contrast between the inner and the outer reservoirs. characteristics of the support of the outer reservoir.
Interplay among the effects of the boundaries just discussed is In Figs. 5 and 6, we consider a dry-gas reservoir and for the
responsible for the changes in flow and production characteristics identification of the conventional flow regimes, we present the
of multiple-fractured horizontal wells in shale reservoirs. results in terms of pseudopressure (Al-Hussainy et al. 1966;
We should also explain our assumption of linear flow in hydrau- Al-Hussainy and Ramey 1966) defined by Eq. A-3 in Appendix A.
lic fractures, which is similar to fractured-vertical-well models. Pseudopressure change is defined as the difference from the initial
This assumption is applicable when the wellbore intercepts the pseudopressure (Δmwf = mi−mwf) and is presented as a function of
hydraulic fracture through its entire thickness. In the case of a time for three different distances to the outer reservoir boundary
fractured horizontal well, the wellbore communicates with the (xe). In Figs. 5 and 6, the case for xe = xF = 250 ft represents the
fracture through a small cylindrical surface piercing through the conditions in which the drainage volume is limited to the inner
fracture surface. As a result of this restricted communication, flow reservoir and the hydraulic-fracture spacing (dF=2ye) is 500 ft. For

250 April 2011 SPE Reservoir Evaluation & Engineering


(a) (b)
A – k f = 2,000 md B – k f = 20,000 md

Fig. 5—Effect of the outer reservoir on multiple-fractured-horizontal-well performance—influence of natural-fracture permeability.

this fracture spacing, the effect of fracture interference (that is, the from the outer reservoir is even more insignificant compared with
effect of the no-flow boundary at ye=dF/2) starts at approximately the volume produced from the inner zone). It is important to note
100 hours. The unit-slope behavior of the responses after 1,000 that mwf = 8.69×108 psi2/cp corresponds to pwf = pi−pwf ≈ 4,000
hours is the same as pseudosteady-state flow behavior. This behav- psi for these examples. Because the initial pressure is pi=4,000
ior is a result of the fact that the outer region does not respond psi, the absolute limit of production is reached at approximately
significantly until the inner zone is depleted to a low pressure. The 10.5 years. Therefore, subject to the constraints of the model con-
deviation of the pseudopressures from the xe = xF case (deviation sidered in our work, for the case presented in Fig. 5a, the effect of
from unit-slope behavior) indicates the start of the support of the the outer boundary will not be felt during the life of the well.
outer reservoir. If, for example, the outer region in Figs. 5 and 6 In Fig. 5b, the natural-fracture permeability is increased to
had permeability comparable to that of the inner region, the sup- kf = 20,000 md, which may not be realistic for most natural
port of the outer reservoir could cause a well-defined half-slope fractures. This case, however, may be used to simulate a hydrau-
behavior (unit slope for pseudosteady state in the inner reservoir lic-fracture network (resulting from dense, complex hydraulic-
and half slope for infinite-acting liner flow in the outer reservoir). fracture growth) with a fracture spacing of 1.25 ft (f = 0.8 f/ft).
For the permeability contrast used in Figs. 5 and 6, the value of As indicated by the comparison of the results in Figs. 5a and 5b,
the late-time slope tends toward one-half before becoming unity at higher natural-fracture permeability causes the support of the outer
late times when the outer reservoir goes into depletion under the reservoir to become evident earlier; however, the effect of the outer
effect of its no-flow boundary at xe. reservoir is still outside the practical life span of the well. These
Fig. 5 shows the effect of natural-fracture permeability. In observations support the common practice of small well spacing
Fig. 5a, for a realistic value of natural-fracture permeability (kf = to efficiently drain tight, unconventional reservoirs with fractured
2,000 md), the support of the outer reservoir becomes evi- horizontal wells.
dent after approximately 100 years for a well spacing of 107 ft Fig. 6 demonstrates the effect of natural-fracture density on the
(xe = 5×106 ft). For smaller well spacing, it takes even longer to feel contribution of the outer reservoir. Fig. 6a is the same as Fig. 5b
the effect of the outer boundary (the relative volume of production having a natural-fracture density of f = 0.8 f/ft. In Fig. 6b, the

(a) (b)
A – ρ f = 0.8 f / ft B – ρ f = 2.0 f / ft

Fig. 6—Effect of the outer reservoir on multiple-fractured-horizontal-well performance—influence of natural-fracture density.

April 2011 SPE Reservoir Evaluation & Engineering 251


PSEUDOPRESSURE DERIVATIVE, dmwf /dln t , psi2/cp
1E+9 1E+9
Hydraulic Fractures Natural Fractures Matrix Hydraulic Fractures Natural Fractures Matrix
mwf, psi2/cp ---------------------------- ------------------------- -------------- ---------------------------- ------------------------- --------------
k = 105 md kf = 2000 md φm = 0.05 k = 105 md kf = 2000 md φ = 0.05
F F m
wF = 10-2 ft hf = 10-3 ft cm = 10-5 psi-1 1E+8 wF = 10-2 ft hf = 10-3 ft cm = 10-5 psi-1
1E+8
φ = 0.38 φ = 0.45 φ F = 0.38 φ f = 0.45
F f

cF = 10-5 psi-1 cf = 10-5 psi-1 cF = 10-5 psi-1 cf = 10-5 psi-1


ρ = 0.8 f / ft Matrix km (md) 1E+7 ρ = 0.8 f / ft
PSEUDOPRESSURE CHANGE,

xF = 250 ft f xF = 250 ft f
1E+7 ---------------------
dF = 500 ft 1E-8 dF = 500 ft
1E-6 Matrix km (md)
1E-4 1E+6
--------------------- qF,sc = 100 Mscf/d
1E-2
1E-8
1E+6 1E-1
1E-6 Reservoir
Reservoir
1E+5 1E-4 -------------------
-------------------
1E-2 h = 250 ft
h = 250 ft
1E-1 T = 150oF
T = 150oF γ = 0.55
1E+5 γ = 0.55
qF,sc = 100 Mscf/d 1E+4 pi = 4000 psi
pi = 4000 psi
μ = 0.0184 cp
WITH CHOKING SKIN μ = 0.0184 cp i
i
xe = 500 ft
WITHOUT CHOKING SKIN xe = 500 ft
1E+4 1E+3
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6
TIME, t, hr TIME, t, hr
(a) (b)

Fig. 7—Effect of matrix permeability on multiple fractured horizontal-well performance—naturally fractured inner reservoir.

fracture density is increased to f = 2 f/ft. Comparing the results cannot be used. For the particular example in Figs. 7 and 8, the
in Figs. 6a and 6b, higher fracture density does not cause a notice- effective, or bulk, permeability of the natural-fracture system is
able change in the time required to feel the effect of the outer k f =kf hft/h=1.6 md. Flow capacity of the natural fractures will be
reservoir; and although at higher fracture densities the contribution discussed in the next section. Here, we note that the validity of
of the outer reservoir is more significant at late times, the effect the preceding discussion requires a large permeability contrast
becomes evident only at impractically late times. The discussion between the matrix and fractures; that is, the production of the
in this section, therefore, supports the common interpretation that matrix must be by virtue of the natural fractures as in the dual-
fractured horizontal wells do not drain the region beyond the tips porosity idealization (Warren and Root 1963; Kazemi 1969; de
of the hydraulic fractures for practical matrix permeabilities. Swaan O. 1976). Most tight gas reservoirs and shale-gas reservoirs
satisfy this requirement and the preceding conclusions can be
Effect of Matrix Permeability generalized for practical applications.
When the matrix permeability decreases from milli- to micro- To demonstrate the difference between the performances of
and nanodarcy range, the reservoir description is changed from multiple-fractured horizontal wells in conventional tight (e.g.,
conventional to unconventional. For unconventional reservoirs, a tight gas) and unconventional tight (e.g., shale-gas) formations,
dense fracture network is required for a sufficient effective perme- we present two cases in Fig. 9. The unconventional tight reservoir
ability in the system. Here, we will first present a discussion of the has a natural-fracture network with matrix and natural-fracture
effect of matrix permeability when the inner reservoir is naturally permeabilities of km = 10−6 md and kf = 2,000 md, respectively (the
fractured, as in unconventional tight formations. Then, we will corresponding effective fracture-network permeability is 1.6 md).
present a comparison with the case in which there are no natural The conventional reservoir is homogeneous and has a permeability
fractures, as in conventional tight-sand applications. of k=10−1 md. Despite the large contrast between the matrix perme-
Fig. 7 shows the pseudopressure change vs. time (Fig. 7a) abilities of the two systems, Fig. 9a shows similar pseudopressure
and log-derivative of pseudopressure vs. time (Fig. 7b) for matrix responses. In fact, Fig. 9b shows that the productivity of the higher-
permeabilities in the range of 10−8 md ≤ km ≤ 10−1 md. In Fig.
7a, pseudopressure responses are also shown for the idealized
case of no choking-skin effect (sc) within hydraulic fractures. 1E-3
Fig. 7a indicates that the choking skin masks the characteristics
PRODUCTIVITY INDEX, J / nF , Mscf - cp / d - psi2

of the early-time responses. The derivative responses in Fig. 7b,


however, are not influenced by the choking skin and display all
early-time flow characteristics. This observation emphasizes the 1E-4
Natural Fractures
importance of derivative analysis for multiple-fractured horizontal -------------------------
kf = 2000 md
wells (especially for the identification of flow characteristics from
hf = 10-3 ft
production data in which derivative analysis may not be possible φ = 0.45
Matrix km (md)
f ---------------------
or may be overlooked). -5
cf = 10 psi -1
1E-5 1E-8
Fig. 7 shows that higher matrix permeabilities reduce the pseu- ρ = 0.8 f / ft 1E-6
f Hydraulic Fractures
dopressure drop at intermediate times. At late times, all responses ----------------------------
1E-4
Reservoir
merge, but this is not caused by the lack of the effect of matrix ------------------- kF = 105 md
permeability. Fig. 8 shows the productivity indices for the same h = 250 ft wF = 10-2 ft
cases considered in Fig. 7 but only with the effect of choking 1E-6 T = 150oF
γ = 0.55 φ = 0.38
F Matrix
skin. Productivity indices are computed by using the procedure pi = 4000 psi cF = 10-5 psi-1 --------------
φ = 0.05
outlined in Appendix B with the drainage area equal to the area μ = 0.0184 cp
i
xF = 250 ft m

of the inner reservoir. Note that the productivity index stabilizes xe = 500 ft dF = 500 ft cm = 10-5 psi-1
when boundary-dominated flow is established at late times. Figs. 1E-7
7 and 8 indicate that the well performance improves as the matrix 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
permeability increases but that matrix permeabilities higher than or TIME, t, hr
equal to 10−4 md do not contribute to the well performance. This
indicates that when the maximum flow capacity of the natural- Fig. 8—Effect of matrix permeability on the productivity of a mul-
fracture system is reached, additional productivity of the matrix tiple-fractured horizontal well—naturally fractured reservoir.

252 April 2011 SPE Reservoir Evaluation & Engineering


1E+9 1E-3
Hydraulic Fractures Natural Fractures Matrix Natural Fractures Matrix Hydraulic Fractures

PRODUCTIVITY INDEX, J / nF , Mscf - cp / d - psi2


---------------------------- ------------------------- -------------- ------------------------- ------------------- ----------------------------

mwf, psi2/cp
kF = 105 md kf = 2000 md km = 10-6 md kf = 2000 md km = 10-6 md kF = 105 md
wF = 10-2 ft hf = 10-3 ft φm = 0.05 hf = 10-3 ft φm = 0.05 wF = 10-2 ft
φ = 0.38 φ = 0.45 c = 10-5 psi-1 φ = 0.45 c = 10-5 psi-1 φ F = 0.38
1E+8 F f m f m
cF = 10-5 psi-1 cf = 10-5 psi-1 cf = 10-5 psi-1 cF = 10-5 psi-1
ρ = 0.8 f / ft ρ = 0.8 f / ft
PSEUDOPRESSURE CHANGE,

xF = 250 ft f f xF = 250 ft
dF = 500 ft dF = 500 ft

1E+7 1E-4
qF,sc = 100 Mscf/d
Reservoir
Reservoir
-------------------
-------------------
h = 250 ft
h = 250 ft
1E+6 T = 150oF
T = 150oF γ = 0.55
γ = 0.55 qF,sc = 100 Mscf/d
pi = 4000 psi
pi = 4000 psi
μ = 0.0184 cp NATURALLY FRACTURED
NATURALLY FRACTURED μ = 0.0184 cp i
i
xe = 500 ft HOMOGENOUS, k = 10-1 md
HOMOGENOUS, k = 10-1 md xe = 500 ft
1E+5 1E-5
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
TIME, t, hr TIME, t, hr
(a) (b)

Fig. 9—Effect of matrix permeability on the change in pseudopressure (a) and productivity (b) of a multiple-fractured horizontal
well—homogeneous and naturally fractured inner reservoirs.

permeability conventional, homogeneous reservoir is less than that A higher density of natural fractures, however, increases the flow
of the unconventional tight reservoir with natural fractures. This capacities of both the natural fractures and the matrix because the
example underlines the importance of natural fractures in uncon- increase in surface area for fluid transfer leads to a more significant
ventional tight formations, such as shale. increase in productivity. From a practical viewpoint, this is a desir-
able feature because increasing the permeability of natural fractures
Effect of Natural-Fracture Permeability and may not be a viable option while increasing the density of natural
Density fractures may be accomplished by designing complex hydraulic
In the preceding section, we emphasized the importance of the fractures in brittle formations (Cramer 2008; Cipolla 2009). It can
flow capacity of natural fractures for production from tight for- also be deduced that creating a complex hydraulic-fracture network
mations. Figs. 10 and 11 indicate that a higher natural-fracture of high permeability (but relatively low density) is not a good option
permeability may not significantly improve the well performance if hydraulic fracturing does not also increase the density of natural
while a higher density of natural fractures may have a significant fractures. This indicates that the experience in brittle shale forma-
effect on productivity. As discussed in the preceding section, this tions, such as the Barnett, may not be extrapolated directly to soft-
is because of the relative flow capacities of the matrix and frac- shale formations such as the Marcellus and the Haynesville.
ture systems. A higher permeability increases the flow capacity
of natural fractures but does not change the flow capacity of the Effect of Hydraulic-Fracture Permeability
matrix because the flow capacity of the matrix is proportional to and Spacing
the surface area. If there is not enough fluid transfer from matrix Determining the optimum number of hydraulic fractures to be
to fracture, a high permeability in the natural fractures does not created along the length of the well is an important practical con-
create additional productivity. cern. A relevant, but more conventional, concern is determining

1E+7 1E-3
Natural Fractures Natural Fracture Matrix Natural Fractures Natural Fracture Matrix
------------------------- Permeability
PRODUCTIVITY INDEX, J / nF , Mscf - cp / psi2

--------------- ------------------------- Permeability ---------------


mwf, psi2/cp

hf = 10-3 ft kf (md) km = 10-6 md hf = 10-3 ft kf (md) km = 10-6 md


φ f = 0.45 --------------------- φ m = 0.05 φ f = 0.45 --------------------- φm = 0.05
2000 2000
cf = 10-5 psi-1 50000 cm = 10-5 psi-1 cf = 10-5 psi-1 50000 cm = 10-5 psi-1
ρ = 0.8 f / ft ρ = 0.8 f / ft
f f
PSEUDOPRESSURE CHANGE,

Reservoir
-------------------
h = 250 ft
1E+6 T = 150oF 1E-4
γ = 0.55 Hydraulic Fractures Hydraulic Fractures Reservoir
pi = 4000 psi ---------------------------- ---------------------------- -------------------
μ = 0.0184 cp kF = 105 md kF = 105 md h = 250 ft
i
T = 150oF
xe = 500 ft wF = 10-2 ft wF = 10-2 ft γ = 0.55
φ F = 0.38 φ F = 0.38 pi = 4000 psi
cF = 10-5 psi-1 c = 10-5 psi-1 μ = 0.0184 cp
F i
xF = 250 ft xF = 250 ft xe = 500 ft
qF, sc = 100 Mscf/d
dF = 500 ft dF = 500 ft
1E+5 1E-5
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4

(a) TIME, t, hr (b) TIME, t, hr


A – PSEUDOPRESSURE VS. TIME B – PRODUCTIVITY INDEX VS. TIM E

Fig. 10—Effect of natural-fracture permeability on the change in pseudopressure (a) and productivity (b) of a multiple-fractured
horizontal well.

April 2011 SPE Reservoir Evaluation & Engineering 253


1E+9 1E-3
Fracture Density Natural Fractures Matrix Matrix Hydraulic Fractures
ρ (f / ft) ------------------------- --------------- --------------- ----------------------------

PRODUCTIVITY INDEX, J / nF , Mscf - cp / psi2


mwf, psi2/cp
f
--------------------- kf = 2000 md km = 10-6 md km = 10-6 md kF = 105 md xF = 250 ft
1.2E-0 hf = 10-3 ft φm = 0.05 φm = 0.05 wF = 10-2 ft dF = 500 ft
8.0E-1
1E+8 φ = 0.45 cm = 10-5 psi-1 cm = 10-5 psi-1 φ F = 0.38
4.0E-1 f
2.0E-1 cf = 10-5 psi-1 cF = 10-5 psi-1
1.2E-1 1E-4
PSEUDOPRESSURE CHANGE,

Fracture Density
Hydraulic Fractures ρ (f / ft)
f
---------------------------- ---------------------
1E+7 kF = 105 md xF = 250 ft 1.2E-0
8.0E-1
wF = 10-2 ft dF = 500 ft
Reservoir 4.0E-1
φ F = 0.38 2.0E-1 Reservoir
-------------------
cF = 10-5 psi-1 h = 250 ft 1E-5 1.2E-1 -------------------
Natural Fractures h = 250 ft
T = 150oF
1E+6 γ = 0.55 ------------------------- T = 150oF
kf = 2000 md γ = 0.55
pi = 4000 psi
μ = 0.0184 cp hf = 10-3 ft pi = 4000 psi
i
φ = 0.45 μ = 0.0184 cp
qF, sc = 100 Mscf/d xe = 500 ft f i
cf = 10-5 psi-1 xe = 500 ft
1E+5 1E-6
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
TIME, t, hr TIME, t, hr
(a) (b)
A – PSEUDOPRESSURE VS. TIME B – PRODUCTIVITY INDEX VS. TIME

Fig. 11—Effect of natural-fracture density on the change in pseudopressure (a) and productivity (b) of a multiple-fractured hori-
zontal well.

the optimum conductivity of hydraulic fractures. Fig. 12 is a density. Therefore, it is desirable to obtain higher dimensionless
demonstration of the conventional concept that increasing hydrau- hydraulic-fracture conductivities by creating larger natural-fracture
lic-fracture conductivity, kFwF, leads to an increase in productivity. densities. It must be noted, however, that higher natural-fracture
The incremental benefit, however, decreases as the conductivity densities also increase the effective natural-fracture permeability
increases. From a fundamental viewpoint, higher productivity and, therefore, decrease the dimensionless hydraulic-fracture con-
is achieved by making the dimensionless conductivity, CFD = ductivity. This leads to a complex design problem for hydraulic
kFwF/(kxF), larger. This concept is relatively straightforward for fractures in tight, naturally fractured formations.
the design of hydraulic fractures in conventional tight formations As shown in Fig. 13a, reducing the hydraulic-fracture spacing
(such as tight gas sands) in which higher conductivities can be significantly influences the magnitudes of the intermediate- and
achieved for low reservoir permeability k. A large dimensionless late-time pseudopressures. On the other hand, Fig. 13b shows
conductivity indicates that the hydraulic fracture can transmit all that hydraulic-fracture spacing does not significantly change flow
the fluid provided by the reservoir. characteristics (it does, however, affect the duration of the flow
For unconventional tight formations, dimensionless hydraulic- regimes). It must be noted that the flow rate of each fracture has
fracture conductivity is defined on the basis of the effective fracture been kept constant while changing the hydraulic-fracture spacing
permeability (k = k f ) and may appear to be relatively large; yet, in Fig. 13. If the total flow rate was kept constant, then the indi-
as discussed in the preceding section, flow capacity of the matrix vidual-fracture rates would have to be reduced when the fracture
may not be sufficient to supply the fracture network to its full spacing decreased. This would yield a separation of the pseudo-
flow capacity because of low matrix permeability or low surface pressure curves at early times (with the smallest pseudopressure
area available for fluid transfer because of low natural-fracture change corresponding with the smaller fracture spacing).

1E+9 1E-3
Hydraulic Fractures Natural Fractures Hydraulic Fractures Reservoir
---------------------------- Natural Fractures ------------------------- ---------------------------- -------------------
PRODUCTIVITY INDEX, J / nF , Mscf - cp / psi2
mwf, psi2/cp

wF = 10-2 ft ------------------------- kf = 2000 md wF = 10-2 ft h = 250 ft


kf = 2000 md
φ F = 0.38 hf = 10-3 ft φ F = 0.38
T = 150oF
hf = 10-3 ft γ = 0.55
cF = 10-5 psi-1 φ = 0.45 c = 10-5 psi-1
1E+8 φ = 0.45 f F pi = 4000 psi
xF = 250 ft Fracture Permeability f
cf = 10-5 psi-1 xF = 250 ft μ = 0.0184 cp
kF (md) cf = 10-5 psi-1 i
dF = 500 ft Matrix ρ = 1.2 f / ft dF = 500 ft xe = 500 ft
PSEUDOPRESSURE CHANGE,

------------------------------ ρ = 1.2 f / ft f
f ---------------
1 x 105 km = 10-6 md
8 x 104
1E+7 φm = 0.05 1E-4
5 x 104
3 x 104 cm = 10-5 psi-1
Reservoir
------------------- Fracture Permeability
h = 250 ft kF (md)
1E+6 T = 150oF Matrix
γ = 0.55 ------------------------------ ---------------
pi = 4000 psi 1 x 105 km = 10-6 md
μ = 0.0184 cp 8 x 104 φm = 0.05
qF, sc = 100 Mscf/d
i
5 x 104
xe = 500 ft cm = 10-5 psi-1
3 x 104
1E+5 1E-5
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
TIME, t, hr TIME, t, hr
(a) (b)
A – PSEUDOPRESSURE VS. TIME B – PRODUCTIVITY INDEX VS. TIME

Fig. 12—Effect of hydraulic-fracture conductivity on the change in pseudopressure (a) and productivity (b) of a multiple-fractured
horizontal well.

254 April 2011 SPE Reservoir Evaluation & Engineering


1E+9 1E+9

PSEUDOPRESSURE DERIVATIVE, dmwf / dlnt, psi2/cp


Fracture Spacing Natural Fractures Matrix Fracture Spacing Natural Fractures Matrix
dF ( ft) ------------------------- --------------- dF ( ft) ------------------------- ---------------

mwf, psi2/cp
--------------------- kf = 2000 md km = 10-6 md --------------------- kf = 2000 md km = 10-6 md
1000 hf = 10-3 ft φm = 0.05 1000 hf = 10-3 ft φ = 0.05
750 1E+8 750 m
1E+8 500 φ = 0.45 cm = 10-5 psi-1 φ f = 0.45 cm = 10-5 psi-1
f 500
250 cf = 10-5 psi-1 250 -5
cf = 10 psi -1

ρ = 0.4 f / ft ρ = 0.4 f / ft
PSEUDOPRESSURE CHANGE,

f f
1E+7
Hydraulic Fractures
Hydraulic Fractures
----------------------------
1E+7 ----------------------------
kF = 105 md
kF = 105 md
wF = 10-2 ft Reservoir 1E+6 wF = 10-2 ft Reservoir
φ = 0.38 ------------------- -------------------
F φ F = 0.38
-5
cF = 10 psi -1 h = 250 ft h = 250 ft
T = 150oF cF = 10-5 psi-1 T = 150oF
1E+6 xF = 250 ft γ = 0.55 γ = 0.55
xF = 250 ft
pi = 4000 psi 1E+5 pi = 4000 psi
μ = 0.0184 cp μ = 0.0184 cp
i i
qf, sc = 100 Mscf/d xe = 500 ft qF, sc = 100 Mscf/d xe = 500 ft

1E+5 1E+4
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E+6
TIME, t, hr TIME, t, hr
(a) (b)
A – PSEUDOPRESSURE VS. TIME B – LOG DERIVATIVE VS. TIME

Fig. 13—Effect of hydraulic-fracture spacing on the pseudopressure responses of a multiple-fractured horizontal well.

The productivity of each fracture is expected to increase when increase. This conclusion supports the field practice of creating
the fracture spacing is decreased. As shown in Fig. 14a, however, complex fractures to increase natural-fracture density and decrease
the incremental gain in productivity with each additional fracture hydraulic-fracture spacing.
decreases. To find the total productivity of the system, the produc-
tivity indices in Fig. 14 must be multiplied by the corresponding Conclusions
number of hydraulic fractures. For example, for a 4,000-ft horizon- In this paper, we investigated the performance characteristics of
tal well, 1,000-ft and 250-ft hydraulic-fracture spacing corresponds fractured horizontal wells and discussed the effects of hydraulic
to two and eight hydraulic fractures, respectively. The total produc- fractures and reservoir properties. The trilinear-flow model pre-
tivity of the well with 1,000-ft fracture spacing is then obtained sented in this work has provided an effective tool for understanding
by multiplying the corresponding individual-fracture productivity and predicting the productivities of fractured horizontal wells.
given in Fig. 14 by two. Similarly, the well productivity for 250- The following conclusions are warranted on the basis of this
ft fracture spacing is obtained by multiplying the corresponding work:
productivity index by eight. This would yield a larger productivity 1. Differences in the performance of fractured horizontal wells in
for the well with eight fractures (250-ft spacing) than for the well conventional and unconventional tight formations are mostly
with two fractures (1,000-ft spacing). determined by the existence of a natural- or induced-fracture
To complete our discussions, in Fig. 14b we present individual- network in the latter.
fracture productivities for two values of natural-fracture density 2. Unlike conventional tight gas reservoirs, relatively high matrix
and two values of hydraulic-fracture spacing. As expected, Fig. 14b permeability and increased hydraulic-fracture conductivity
indicates that increasing natural-fracture density while decreas- are not sufficient to have good productivity in unconventional,
ing hydraulic-fracture spacing provides the greatest productivity shale-gas reservoirs. Existence of a dense natural- or induced-

1E-3 1E-3
Fracture Spacing Natural Fractures Matrix Natural Fractures Hydraulic Fractures Matrix Reservoir
PRODUCTIVITY INDEX, J / nF , Mscf - cp / d - psi2

------------------------- ---------------------------- --------------- -------------------


PRODUCTIVITY INDEX, J / nF , Mscf - cp / psi2

dF ( ft) ------------------------- ---------------


kf = 2000 md km = 10-6 md kf = 2000 md kF = 105 md km = 10-6 md h = 250 ft
---------------------
hf = 10-3 ft T = 150oF
1000 hf = 10-3 ft φm = 0.05 wF = 10-2 ft φm = 0.05 γ = 0.55
750 φ = 0.45 φ = 0.45 φ = 0.38 c = 10-5 psi-1
500 f cm = 10-5 psi-1 f F m pi = 4000 psi
250 cf = 10-5 psi-1 cf = 10-5 psi-1 cF = 10-5 psi-1 μ = 0.0184 cp
Reservoir i
ρ = 0.4 f / ft ------------------- xF = 250 ft xe = 500 ft
f
h = 250 ft
T = 150oF
1E-4 γ = 0.55 1E-4
pi = 4000 psi
Hydraulic Fractures μ = 0.0184 cp
i
----------------------------
xe = 500 ft
kF = 105 md
wF = 10-2 ft
φ = 0.38 Natural Fracture Fracture Spacing
F
Density ρf (f / ft) dF ( ft)
cF = 10-5 psi-1
------------------------ ---------------------
xF = 250 ft 0.4 1000
1.2 250
1E-5 1E-5
1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1E-3 1E-2 1E-1 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5

(a) TIME, t, hr (b) TIME, t, hr

A – EFFECT OF HYDRAULIC FRACTURE SPACING FOR B – COMBINED EFFECT OF HYDRAULIC FRACTURE


FIXED NATURAL FRACTURE DENSITY SPACING AND NATURAL FRACTURE DENSITY

Fig. 14—Effect of hydraulic-fracture spacing and natural-fracture density on the productivity of multiple-fractured horizontal
wells.

April 2011 SPE Reservoir Evaluation & Engineering 255


fracture network is a more important requirement for favorable mwf = flowing pseudopressure, psi2/cp
productivities in shale-gas reservoirs. However, increasing the nf = number of natural fractures
permeability of the fracture network has a relatively less sig- nF = number of hydraulic fractures
nificant effect on productivity than increasing the density of the p = pressure, psi
fracture network. pb = base pressure, psi
3. The limitations of the interaction between the tight matrix and
pi = initial reservoir pressure, psi
natural-fracture network may place restrictions on productivity
improvements. pwD = dimensionless flowing wellbore pressure
4. Even for relatively large matrix permeabilities, the drainage pwf = flowing wellbore pressure, psi
volume of fractured horizontal wells is limited to the naturally p = average reservoir pressure, psi
or artificially fractured inner reservoir between the hydraulic p D = dimensionless average reservoir pressure
fractures. Under these conditions, flow convergence is linear pwD = dimensionless flowing wellbore pressure in the Laplace
for most of the productive life of the well. space
5. Hydraulic-fracture conductivities should be optimized on the pwD,storage = dimensionless wellbore-storage pressure in the Laplace
basis of the flow capacities of the matrix and natural-fracture space
network. Decreasing fracture spacing increases the productivity q = production rate, B/D (oil), Mcf/D (gas)
of the well, but the incremental gain for each additional fracture qF = individual-hydraulic-fracture production rate, B/D (oil),
decreases. scf/D (gas)
qF,sc = individual-hydraulic-fracture production rate at surface
Nomenclature conditions, STB/D (oil), Mscf/D (gas)
A = drainage area, ft2 qsc = production rate at surface conditions, STB/D (oil),
AD = dimensionless drainage area Mscf/D (gas)
B = formation volume factor, bbl/STB rw,eff = effective wellbore radius, ft
ct = total compressibility, psi−1 rwj = fracture effective wellbore radius, ft
ctf = total natural-fracture compressibility, psi−1 s = Laplace-transform parameter
ctF = total hydraulic-fracture compressibility, psi−1 sc = hydraulic-fracture flow-convergence skin factor (Eq.
ctI = total compressibility of the inner reservoir, psi−1 A-23)
ctIi = initial total compressibility of the inner reservoir, t = time, hours
psi−1 tD = dimensionless time
ctm = total matrix compressibility, psi−1 T = reservoir temperature, °R
ctO = total compressibility of the outer reservoir, psi−1 u = parameter defined as the product of the dual-porosity
CA = shape factor function and the Laplace-transform parameter (Eq.
CD = dimensionless wellbore-storage coefficient A-18)
CF = hydraulic-fracture conductivity, md-ft wF = hydraulic-fracture width, ft
CF,eff = effective hydraulic-fracture conductivity, md-ft wF,eff = effective hydraulic-fracture width, ft
CFD = dimensionless hydraulic-fracture conductivity wFD = dimensionless hydraulic-fracture width
CFD,eff = effective dimensionless hydraulic-fracture conductivity xF = hydraulic fracture half-length, ft
CRD = dimensionless reservoir conductivity xF,eff = effective hydraulic-fracture half-length, ft
dF = distance between two hydraulic fractures, ft xe = reservoir size in x-direction, ft
D = distance between outermost hydraulic fractures, ft xeD = dimensionless reservoir size in x-direction
f = dual-porosity function defining fluid transfer from ye = reservoir size in y-direction, ft
matrix to fracture (Eq. A-20) ye,eff = effective reservoir size in y-direction, ft
h = reservoir thickness, ft yeD = dimensionless reservoir size in y-direction
hf = thickness of natural fractures, ft z = real-gas compressibility factor
hft = total thickness of natural fractures, ft F = inner-reservoir parameter used in trilinear-flow model
hF = height of hydraulic fractures, ft (Eq. A-10)
hm = thickness of matrix slabs, ft R = outer-reservoir parameter used in trilinear-flow model
J = transient productivity index, STB/D/psi (oil), Mscf/D/ (Eq. A-13)
psi2/cp (gas) F = inner-reservoir parameter used in trilinear-flow model
k = homogeneous reservoir permeability, md (Eq. A-12)
kf = natural-fracture permeability, md R = outer-reservoir parameter used in trilinear-flow model
kF = hydraulic-fracture permeability, md (Eq. A-15)
kI = permeability of the inner reservoir, md  = Euler’s constant (= 0.5772…)
km = matrix permeability (dual-porosity model), md I = inner-reservoir diffusivity, ft2/hr
kO = permeability of the outer reservoir, md f = natural-fracture diffusivity, ft2/hr
k f = effective (bulk) natural-fracture permeability, md F = hydraulic-fracture diffusivity, ft2/hr
kI = effective (bulk) permeability of the inner reservoir, FD = hydraulic-fracture diffusivity ratio
md O = outer-reservoir diffusivity, ft2/hr
m = pseudopressure (Eq. A-2), psi2/cp RD = outer-reservoir diffusivity ratio
mi = initial pseudopressure, psi2/cp
= dual-porosity transmissivity ratio (Eq. A-22)
mwD = dimensionless flowing pseudopressure  = fluid viscosity, cp
m = average reservoir pseudopressure, psi2/cp i = initial fluid viscosity, cp
m D = dimensionless average pseudopressure f = density of natural fractures, fractures/ft
mwD = dimensionless flowing wellbore pseudopressure in the  = reservoir porosity, fraction
Laplace space f = natural-fracture porosity, fraction
mwD,storage = dimensionless wellbore-storage pseudopressure in the F = hydraulic-fracture porosity, fraction
Laplace space I = inner-reservoir porosity, fraction

256 April 2011 SPE Reservoir Evaluation & Engineering


Ii = initial inner reservoir porosity, fraction Larsen, L. and Hegre, T.M. 1994. Pressure Transient Analysis of Multifrac-
m = matrix porosity, fraction tured Horizontal Wells. Paper SPE 28389 presented at the SPE Annual
O = outer-reservoir porosity, fraction Technical Conference and Exhibition, New Orleans, 25–28 September.
= dual-porosity storativity ratio (Eq. A-21) doi: 10.2118/28389-MS.
Maxwell, S.C., Waltman, C.K., Warpinski, N.R., Mayerhofer, M.J., and
Acknowledgments Boroumand, N. 2009. Imaging Seismic Deformation Induced by
Hydraulic Fracture Complexity. SPE Res Eval & Eng 12 (1): 48–52.
This research has been conducted under the Marathon Center of SPE-102801-PA. doi: 10.2118/102801-PA.
Excellence for Reservoir Studies (MCERS). Parts of this work Mayerhofer, M.J., Bolander, J.L., Williams, L.I., Pavy, A., and Wolhart,
have been completed to fulfill the MS degree requirements of S.L. 2005. Integration of Microseismic-Fracture-Mapping Fracture and
Margaret Brown at the Colorado School of Mines. The financial Production Analysis With Well Interference Data to Optimize Fracture
support for Margaret Brown’s graduate studies has been provided Treatments in the Overton Field, East Texas. Paper SPE 95508 pre-
by the MCERS and the Petroleum Engineering Department at the sented at the SPE Annual Technical Conference and Exhibition, Dallas,
Colorado School of Mines (CSM). 9–12 October. doi: 10.2118/95508-MS.
Medeiros, F., Ozkan, E., and Kazemi, H. 2008. Productivity and Drainage
References Area of Fracture Horizontal Wells in Tight Gas Reservoirs. SPE Res
Al-Hussainy, R. and Ramey, H.J. Jr. 1966. Application of Real Gas Flow Eval & Eng 11 (5): 902–911. SPE-108110-PA. doi: 10.2118/108110-
Theory to Well Testing and Deliverability Forecasting. J Pet Technol PA.
18 (5): 637–642; Trans., AIME, 237. SPE-1243-B-PA. doi: Mukherjee, H. and Economides, M.J. 1991. A Parametric Comparison
10.2118/1243-B-PA. of Horizontal and Vertical Well Performance. SPE Form Eval 6 (2):
Al-Hussainy, R., Ramey, H.J. Jr., and Crawford, P.B. 1966. The Flow of 209–216. SPE-18303-PA. doi: 10.2118/18303-PA.
Real Gases Through Porous Media. J Pet Technol 18 (5): 624–636; Prats, M. 1961. Effect of Vertical Fractures on Reservoir Behavior—Incom-
Trans., AIME, 237. SPE-1243-A-PA. doi: 10.2118/1243-A-PA. pressible Fluid Case. SPE J. 1 (2): 105–118; Trans., AIME, 222. SPE-
Al-Kobaisi, M., Ozkan, E., and Kazemi, H. 2006. A Hybrid Numerical/ 1575-G. doi: 10.2118/1575-G.
Analytical Model of a Finite-Conductivity Vertical Fracture Intercepted Raghavan, R.S., Chen, C.C., and Agarwal, B. 1997. An Analysis of Hori-
by a Horizontal Well. SPE Res Eval & Eng 24 (10): 345–355. SPE- zontal Wells Intercepted by Multiple Fractures. SPE J. 2 (3): 235–245.
92040-PA. doi: 10.2118/92040-PA. SPE-27652-PA. doi: 10.2118/27652-PA.
Bennett, C.O., Camacho-V., R.G., Reynolds, A.C., and Raghavan, R. 1985. Serra, K.V., Reynolds, A.C., and Raghavan, R. 1983. New Pressure Tran-
Approximate Solutions for Fractured Wells Producing Layered Reser- sient Analysis Method for Naturally Fractured Reservoirs. J Pet Technol
voirs. SPE J. 25 (5): 729–742. SPE-11599-PA. doi: 10.2118/11599-PA. 35 (12): 2271−2283. SPE-10780-PA. doi: 10.2118/10780-PA.
Brown, M. 2009. Analytical Trilinear Pressure Transient Model for Mul- Soliman, M.Y., Hunt, J.L., and El Rabaa, A.M. 1990. Fracturing Aspects
tiply Fractured Horizontal Wells in Tight Shale Reservoirs. MS thesis, of Horizontal Wells. J Pet Technol 42 (8): 966–973. SPE-18542-PA.
Colorado School of Mines, Golden, Colorado, USA. doi: 10.2118/18542-PA.
Brown, M., Ozkan, E., Ragahavan, R., and Kazemi, H. 2009. Practical Warren, J.E. and Root, P.J. 1963. The Behavior of Naturally Fractured
Solutions for Pressure Transient Responses of Fractured Horizontal Reservoirs. SPE J. 3 (3): 245–255; Trans., AIME, 228. SPE-426-PA.
Wells in Unconventional Reservoirs. Paper SPE 125043 presented at doi: 10.2118/426-PA.
SPE Annual Technical Conference and Exhibition, New Orleans, 4–7
October. doi: 10.2118/125043-MS. Appendix A—Trilinear-Flow Model:
Camacho-V., R.G. 1984. Response of Wells Producing Commingled Res-
Analytical Expressions and Definitions
ervoirs: Unequal Fracture Length. MS thesis, University of Tulsa,
Tulsa. Here, we present the analytical expressions for the trilinear model
Camacho-V., R.G., Raghavan, R. and Reynolds, A.C. 1987. Response of used in this paper. For a detailed derivation of these expressions,
Wells Producing Layered Reservoirs: Unequal Fracture Length. SPE we refer the reader to Brown et al. (2009) and Brown (2009). For
Form Eval 2 (1): 9–28. SPE-12844-PA. doi: 10.2118/12844-PA. convenience, we introduce the trilinear-flow model in terms of
Chen, C.-C. and Raghavan, R. 1997. A Multiply-Fractured Horizontal Well dimensionless variables. We define a dimensionless pressure by
in a Rectangular Drainage Region. SPE J. 2 (4): 455–465. SPE-37072-
PA. doi: 10.2118/37072-PA. pwD =
kI h
141.2qF ,sc B
( )
pi − pwf , . . . . . . . . . . . . . . . . . . . . . (A-1)
Cinco-Ley, H. and Samaniego-V., F. 1981. Transient Pressure Analysis for
Fractured Wells. J Pet Technol 33 (9): 1749-1766. SPE-7490-PA. doi:
10.2118/7490-PA. for liquid flow, and
Cipolla, C.L. 2009. Modeling Production and Evaluating Fracture Perfor-
mance in Unconventional Gas Reservoirs. J Pet Technol 61 (9): 84–90. mwD =
kI h
1422qF ,scT ⎣ ⎦ ( )
⎡ m ( pi ) − m pwf ⎤, . . . . . . . . . . . . . . . . (A-2)
SPE-118536-MS. doi: 10.2118/118536-MS.
Cramer, D.D. 2008. Stimulating Unconventional Reservoirs: Lessons
Learned, Successful Practices, Areas for Improvement. Paper SPE for gas flow, with m(p) denoting real-gas pseudopressure (Al-Hus-
114172 presented at the SPE Unconventional Reservoirs Conference, sainy et al. 1966; Al-Hussainy and Ramey 1966),
Keystone, Colorado, USA, 10–12 February. doi: 10.2118/114172-MS. p
p′
de Swaan O., A. 1976. Analytical Solutions for Determining Naturally m ( p) = 2 ∫ dp′ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-3)
Fractured Reservoir Properties by Well Testing. SPE J. 16 (3): 117-122; pb
z
Trans., AIME, 261. SPE-5346-PA. doi: 10.2118/5346-PA.
Fisher, M.K., Heinze, J.R., Harris, C.D., Davidson, B.M., Wright, C.A., We note that kI h = kf hft for a naturally fractured inner zone
and Dunn, K.P. 2004. Optimizing Horizontal Completion Techniques Dimensionless time is defined by
in the Barnett Shale Using Microseismic Fracture Mapping. Paper SPE
2.637 × 10 −4 kI t
90051 presented at the SPE Annual Technical Conference and Exhibi- tD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-4)
tion, Houston, 26–29 September. doi: 10.2118/90051-MS. ⎡⎣(ct )I ⎤⎦ i x F2
i
Gringarten, A.C., Ramey, H.J. Jr., and Raghavan, R. 1974. Unsteady-State
Pressure Distributions Created by a Well With a Single Infinite-Con- where t is in hours and (ct)I = (ct)f for a naturally fractured inner
ductivity Vertical Fracture. SPE J. 14 (4): 347–360; Trans., AIME, 257. zone. The dimensionless variables of the reservoir and fracture
SPE-4051-PA. doi: 10.2118/4051-PA. geometry are defined by
Kazemi, H. 1969. Pressure Transient Analysis of Naturally Fractured
xe
Reservoir With Uniform Fracture Distribution. SPE J. 9 (4): 451–462; xeD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-5)
Trans., AIME, 246. SPE-2156-A. doi: 10.2118/2156-PA. xF

April 2011 SPE Reservoir Evaluation & Engineering 257


ye for a naturally fractured inner reservoir with the transient dual-
yeD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-6) porosity idealization (Serra et al. 1983). In Eq. A-20, the dual-
xF
porosity parameters are defined by
and
(ct h )m
= . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-21)
wFD
w
= F. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-7) (ct h ) f
xF
and
The dimensionless trilinear-flow solution for a multiple-fractured
horizontal well is given, in the Laplace-transform domain, by 12 x F2 km

= . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-22)
hm h f k f

pwD = mwD = + sc, . . . . . . . . . . . . . . (A-8)
CFD s  F tanh ( F ) All properties appearing in Eqs. A-21 and A-22 are intrinsic.
In Eq. A-8, sc denotes the contribution of flow choking around
where s is the Laplace-transform parameter with respect to the the intersection of the horizontal well and hydraulic fracture.
dimensionless time, tD, defined in Eq. A-4, Mukherjee and Economides (1991) have provided the following
expression to estimate sc:
k F wF
CFD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-9) kI h ⎡ ⎛ h ⎞ ⎤
kI x F sc = ⎢ ln − ⎥ . . . . . . . . . . . . . . . . . . . . . . . . (A-23)
kF wF ⎣ ⎜⎝ 2rw ⎟⎠ 2 ⎦

2F s The expression given by Eq. A-23 should be valid after the end of
F = + , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-10) the radial flow within the hydraulic fracture, which corresponds to
CFD FD
a very short period of time at the beginning of flow. For practical
cases, the early-time radial flow within the hydraulic fracture is
F completely masked by the effect of wellbore storage. The effect
FD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-11) of wellbore storage can be incorporated easily by using the fol-
I
lowing relationship:

pwD
F =  R tanh ⎡⎣  R ( yeD − wFD 2 ) ⎤⎦ , . . . . . . . . . . . . . . . (A-12) pwD ,storage = mwD ,storage = , . . . . . . . . . . . . . . . . (A-24)
1 + C D s 2 pwD

where pwD is the trilinear-model solution given by Eq. A-8 and CD


R is the dimensionless wellbore-storage coefficient given by
R = + u, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-13)
C RD yeD
 cC
CD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-25)
2 (ct h )I x F2
kI x F
C RD = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-14) where c = 5.615 or 1 for oil and gas flow, respectively.
kO ye

R = s RD tanh ⎡⎣ s RD ( xeD − 1) ⎤⎦ , . . . . . . . . . . . . . . (A-15) Appendix B—Computation of the


Productivity Index
The productivity index of a multiple-fractured horizontal well is
and defined by
O
RD = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-16) J=
qsc q
= nF F ,sc , . . . . . . . . . . . . . . . . . . . . . . . . . (B-1)
I p − pwf p − pwf
In Eqs. A-11 and A-16, ( = F , I , or O ) is defined by
for liquid flow, where p denotes the average pressure in the reser-
k voir. We rearrange Eq. B-1 to write
= . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-17)
(ct )  qF ,sc
J = nF , . . . . . . . . . . . . . . . . . . . (B-2)
141.2qF ,sc B
For a naturally fractured inner reservoir, I = f in Eqs. A-9, A-11,
A-14, A-16, and A-17. Also, for a naturally fractured inner zone, kI h
( pwD − p D )
all definitions are based on intrinsic properties, except for kI = k f where pwD is computed from Eq. A-8 and
in Eqs. A-9 and A-14 where k f is the bulk (effective) permeability
(k f = k f h ft h ). In Eq. A-13, tD
p D = 2 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-3)
(1 + ) AD
u = sf ( s ), . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-18)
with
where
A
AD = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-4)
x F2
f ( s ) = 1, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-19)
For gas flow, we define
for a homogeneous inner reservoir and
qsc qF ,sc
J= = nF , . . . . . . . . . . . . (B-5)

⎛ 3 s ⎞ m − mwf 1422qF ,scT
f (s) = 1 + tanh ⎜ ⎟ , . . . . . . . . . . . . . . . . . . . . (A-20)
( mwD − m D )
3s ⎝
⎠ kI h

258 April 2011 SPE Reservoir Evaluation & Engineering


where m denotes the average pseudopressure in the reservoir and 2007 SPE Formation Evaluation Award and a Distinguished
m D can be approximated by Member of SPE. Margaret Brown is a reservoir engineer work-
ing for Marathon Oil in Houston. E-mail: mbrown6@marathonoil.
tD com. She holds an MSc degree in petroleum engineering from
m D ≈ 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-6) CSM. Rajagopal Raghavan retired from Phillips Petroleum as its
(1 + ) AD formation evaluation manager in Bartlesville, Oklahoma, USA.
His interests are in well performance and reservoir description
Although we use constant-rate-production equations in this paper, methods. Raghavan has served SPE in various capacities over
the productivity index can also be used for variable-rate condi- the years, including serving as Senior Technical Editor, and is
tions. A more detailed discussion around computation of the gas a Distinguished and an Honorary Member of SPE. Hossein
productivity index (including variable-rate production) is given by Kazemi is the Chesebro’ Distinguished Professor of Petroleum
Medeiros et al. (2008). Engineering at CSM and Codirector of MCERS. E-mail: hka-
zemi@mines.edu. He holds BS and PhD degrees in petroleum
Erdal Ozkan is a professor of petroleum engineering and co- engineering from the University of Texas at Austin (1961 and
director of MCERS at CSM. E-mail: eozkan@mines.edu. He holds 1963, respectively). Kazemi is a Distinguished and an Honorary
BS and MS degrees from Istanbul Technical University and a Member of SPE and has served as Distinguished Author and
PhD degree from the University of Tulsa, all in petroleum engi- Speaker for SPE. He retired from Marathon in 2001 after serving
neering. Ozkan’s main research interests are pressure-tran- as senior technical consultant, director of production research,
sient analysis, modeling fluid flow in porous media, horizontal manager of reservoir technology, and executive technical
and multilateral well technology, and unconventional reser- fellow at Marathon Petroleum Technology Center in Littleton,
voirs. He has served as the Executive Editor of SPE Res Eval & Colorado, USA. At CSM, Kazemi has taught graduate courses
Eng and chaired several SPE forums, workshops, and confer- in petroleum engineering and supervised graduate research
ences. Ozkan is a member of the SPE Reservoir Description in reservoir modeling, well testing, and improved oil and gas
and Dynamics Advisory Committee. He is the recipient of the recovery.

April 2011 SPE Reservoir Evaluation & Engineering 259


Gas Permeability of Shale
A. Sakhaee-Pour and Steven L. Bryant, The University of Texas at Austin

Summary within organic material, methane molecules can adsorb to an


Permeability is one of the most fundamental properties of any res- appreciable extent. For instance, the influence of the adsorbed
ervoir rock required for modeling hydrocarbon production. How- layer on the void space was considered and the adsorbed porosity
ever, shale permeability has not yet been understood fully because was introduced (Cui et al. 2009). In addition, the volume of
of the complexities involved in modeling flow through nanoscale adsorbed gas was estimated using reservoir simulation (Cipolla
throats. In this paper, we analyze the effects of adsorbed layers of et al. 2010). The volume and producibility of adsorbed methane
methane (CH4) and of gas slippage at pore walls on the flow have direct implications for reserves and, thus, have been widely
behavior in individual conduits of simple geometry and in net- discussed. However, the impact of an adsorbed layer on perme-
works of such conduits. The network is based on scanning elecron ability has been less widely explored. Yet because the voids are
microscope (SEM) images and a drainage experiment in shale. To so small, the adsorbed layer can also affect the transport proper-
represent the effect of adsorbed gas, the effective size of each ties of the shale.
throat in the network depends on the pressure. The hydraulic con- This paper analyzes the dependency of the in-situ shale-matrix
ductance of each throat is determined on the basis of the Knudsen- permeability on reservoir pressure. In this regard, we implement
number (Kn) criterion (Knudsen 1909). The combined effects of the effect of an adsorbed layer in simple conduit geometries as a
adsorption and slip depend strongly on pressure and on conduit di- function of gas pressure. Subsequently, we adopt noncontinuum
ameter. The results indicate that laboratory measurements made flow models to account for nonzero slip velocity in the conduits.
with N2 at ambient temperature and 5-MPa pressure, which is typi- The presence of an adsorbed layer of gas molecules has a geomet-
cal for a transient pulse-decay (TPD) method, overestimate the gas ric implication (it reduces the cross section available for transport)
permeability at early life of production by a factor of four. This ra- and an influence on microscale boundary condition (slippage of
tio increases if the measurement is run at ambient condition gas being transported through the pore). Here, we examine a mac-
because the low pressure enhances the slippage and reduces the roscopic consequence of these phenomena—namely, the effect on
thickness of the adsorbed layer. Moreover, the permeability the permeability of the rock to gas—by implementing these phe-
increases nonlinearly as the in-situ pressure decreases during pro- nomena in a network model of the void space. The effect of
duction. This effect contributes to mitigating the decline in the pro- adsorption and slippage depends on the size of the conduit, and
duction rate of shale-gas wells. Laboratory data available in the though a network is not a literal model of shale pore space, it does
literature for CH4 permeability at pressures below 7 MPa agree serve as a convenient vehicle for examining the effect of a distri-
with model predictions of the effect of pressure. bution of pore sizes. The results have implications for interpreting
measurements of shale permeability to gas at laboratory condi-
Introduction tions and for interpreting flow rates from production wells over
Shale has provoked a great deal of research recently because of time. The flow behavior from a rock mass into a production well
the considerable volume of natural gas stored in reservoirs. The is controlled by resistance of flow paths that are combinations of
natural gas produced from these reservoirs is called shale gas and matrices and fractures. Thus, for operational decisions, the effect
is envisaged to provide a substantial fraction of US gas produc- of pressure on matrix permeability should be incorporated with
tion, with some reports of as much as half by 2020 (Polczer natural- and induced-fracture dependency on the pressure (Philip
2009). However, debate continues about how much gas can be et al. 2005).
produced from these types of reservoirs (Urbina 2011). We refer to laboratory and field conditions extensively in the
An accurate measurement of shale permeability is challenging following discussion, and for the sake of clarity, we define them
because it is so small, on the order of 10–21 m2 (nanodarcies). The here. The temperature is equal to 300 and 360 K at the laboratory
challenge appears in both experimental and theoretical investiga- and field conditions, respectively. The pressure at the laboratory
tions. For instance, constant-pressure steady-state-flow measure- condition is presumed to be 5 MPa, which is a common pressure
ment is not used for shale because it requires a considerable time for permeability measurement using TPD (Billiote et al. 2008).
(Mallon and Swarbrick 2008). Consequently, other approaches The gas permeability of shale at this condition is on the order of
such as TPD (Billiotte et al. 2008; Cui et al. 2009) and crushed 10 nanodarcies. We assume the thickness of the adsorbed layer is
rock (Luffel et al. 1993) are commonly employed. These methods negligible at laboratory condition. This is a plausible assumption
reduce the time required for measurement by changing the bound- because the ratio of no-slip hydraulic conductance without an
ary condition and sample size. The test is altered from steady state adsorbed layer to no-slip hydraulic conductance with an adsorbed
to transient in TPD by changing the constant-pressure boundary layer at 5 MPa is close to unity, regardless of throat size, as we
conditions to pressures that vary with time. Therefore, it is con- will show later in the paper. For estimating the gas conductance
ducted more quickly because sustained flow at steady state is not and permeability, we further suppose that gases at the laboratory
attempted. In the crushed-rock method, the rock is broken into and in-situ conditions are N2 and CH4, respectively. For conven-
small pieces so the transient flow of gas out of the pieces is more ience, we use the terms “liquid conductance” and “liquid perme-
rapid. ability” to refer to hydraulic conductance computed with a
Studies of gas flow through shale have invoked a different set continuum model and no-slip boundary condition at pore walls.
of parameters. For example, the gas evacuated from the core was We use the terms “gas conductance” and “gas permeability” to
modeled in terms of conventional permeability and desorption refer to the conductance when slippage is taken into account.
(Javadpour et al. 2007). Because many of the voids in shale are Both “gas” and “liquid” conductance can be computed when an
adsorbed layer is present. We denote the gas and liquid perme-
abilities at the laboratory condition by kg,lab and kl,lab, respec-
tively. The pressure varies at the field condition, controlling the
Copyright V
C 2012 Society of Petroleum Engineers
thickness of the adsorbed layer. The pressure also affects the Kn
This paper (SPE 146944) was accepted for presentation at the SPE Annual Technical and, hence, the slippage. For the field condition, the gas and liq-
Conference and Exhibition, Denver, 30 October–2 November 2011, and revised for
publication. Original manuscript received for review 28 June 2011. Revised manuscript
uid permeabilities are represented by kg,in-situ and kl,in-situ,
received for review 17 January 2012. Paper peer approved 29 March 2012. respectively.

August 2012 SPE Reservoir Evaluation & Engineering 401


kl
TABLE 1—FLUID-FLOW REGIMES DEFINED BY Jtotal ¼ Jvisc þ JKn ¼  nrp  DKn rn
l
RANGES OF Kn  
kl rp
¼ p þ DKn ;
Kn 0–10–3 10–3–10–1 10–1–101 >101 l RT                   ð4Þ
Flow regime Continuum Slip Transition Free-molecular where rp is the pressure gradient and DKn is the Knudsen diffu-
sivity coefficient. This relation can also be written in terms of gas
permeability, as
 
Gas-Flow Regimes kl rp kg p rp
Jtotal ¼  p þ DKn ¼ ; . . . . . . . . . . . ð5Þ
Kn (Knudsen 1909) is used to differentiate flow regimes in con- l RT l RT
duits at micro- and nanoscale. This nondimensional parameter is
where kg is the single-phase gas permeability. By substituting the
defined as
Knudsen diffusivity coefficient and implementing the ideal-gas
k assumption, the gas permeability relative to the no-slip permeabil-
Kn ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ ity in terms of Kn is yielded
K
 
where k denotes the mean free path of gas molecules and K the kg DK l
¼ 1þ n ¼ 1 þ k1 Kn ; . . . . . . . . . . . . . . . . . ð6Þ
characteristic length of the channel. The flow regime changes kl kl p
from continuum model to discrete particles as the Kn increases.
Substituting the mean free path yields the following relation (Roy where k1 is the permeability enhancement equal to 13.58 for the
et al. 2003): circular cross section, as clarified in Appendix A. This relation
rffiffiffiffiffiffiffiffiffi indicates that the first-order models (Eq. 3) and the DGM (Eq. 6)
p l suggest a similar trend for the gas permeability (i.e., a linear func-
Kn ¼  ; . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ tion of Kn).
2RT qK
Eq. 6 is the analog of Klinkenberg’s correction (Klinkenberg
where l is the gas viscosity, q the density, T the prevailing tem- 1941) for a single conduit. In the Klinkenberg correction, the perme-
perature, and R the specific gas constant. The specific gas constant ability to gas is evaluated at several pressures for a core sample
is the ratio of universal gas constant R to the molar mass m. On instead of a single conduit. For steady-flow experiments, the pres-
the other hand, the length scale K is equal to the diameter and sure is replaced by the average of the inlet and outlet pressures across
aperture height in the circular tube and slit, respectively. For N2 the core. The term DKknl l is determined empirically as the slope of the
and CH4 at the range of temperatures (300 to 360 K) and pres- measured gas permeabilities plotted against the reciprocal pressure.
sures (5 to 28 MPa) of interest here, the range of values of k is 0.3 While the first-order slip model and the DGM indicate a quali-
to 2.2 nm. For pore sizes K in the range of 3 to 30 nm, we thus en- tatively similar trend for the gas permeability, the rate of change
counter values for Kn in the range of 10–2 to 100. with Kn is different. The high Kn is attained by decreasing the
throat size in the first-order slip model. In this case, molecule/mol-
ecule collisions are not necessarily infrequent; rather, they are less
Continuum Regime. The flow regime is categorized on the basis frequent than molecule/wall collisions. On the other hand, the low
of Kn, as summarized in Table 1 (Roy et al. 2003). This table indi- density of the gas is the reason for the slippage in the DGM. Note
cates that the Navier-Stokes equations with no-slip boundary con- that reductions in both the throat size and the density increase Kn
dition are appropriate only for small Kn (Kn < 10–3). In other (see Eq. 2). Consequently, the permeability enhancements are
words, the classical continuum-model assumptions (momentum obtained by measuring the flow rates. The in-situ condition for gas
transfer by means of bulk phase viscosity, fluid velocity matches shale corresponds to the large gas density (approximately 100 kg/
solid velocity at walls) cannot always provide accurate results. m3). Kn in this situation is large because shale conduits are very
small. Thus, in this study, we use the first-order slip model in the
Slip Flow Regime. The ratio of molecule/wall to molecule/mole- range of 103< Kn < 101.
cule collisions increases as Kn increases. However, the molecule/
molecule interaction is still dominant in the slip flow regime. Transition Flow Regime. The physics of the transition flow re-
Therefore, Navier-Stokes equations were suggested (Karniadakis gime is complicated, and most models are used to predict the
et al. 2005) to remain valid for this domain (103< Kn < 101). computational results of the Monte Carlo simulations (Karniada-
A first-order slip-boundary condition was applied to include the kis et al. 2005). The proposed models (Karniadakis et al. 2005)
effect of molecule/wall collisions (Roy et al. 2003). After imple- implement different shear stress laws in the Navier-Stokes equa-
menting the tangential momentum coefficient as shown in Appen- tions. The higher-order gradients of the velocity are implemented
dix A, the gas permeability can be estimated as in the proposed models, and the velocity profile is subject to
higher-order boundary conditions. To capture the flow rate, the
kg ¼ kl ð1 þ a1 Kn Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ corresponding mass-flow rate for a circular tube was proposed as
(Karniadakis et al. 2005)
where kl is the permeability of the conduit to liquid (when the no-  
slip boundary condition applies), kg the gas permeability of the   4Kn pR4
conduit, and a1 the permeability enhancement. The liquid perme- M_ ¼ 1 þ 1:358tan1 ð4Kn0:4 Þ  Kn 1 þ qrp;
1 þ Kn 8l
ability is assumed constant (i.e., independent of pressure) and esti-
mated on the basis of the characteristic length size of the throat                    ð7Þ
from conventional theory. The coefficient a1 depends on the ge- where M_ is the mass flux and q is the average density. The coeffi-
ometry of the conduit. It is approximately 5 for circular cross sec- cients are obtained by fitting the equation to the Monte Carlo sim-
tions (see Appendix A). ulation results. We can express the apparent permeability by
The slip flow regime can also be simulated using the dusty gas dividing the mass-flow rate to the no-slip condition. Therefore,
model (DGM) (Mason and Malinauskas 1983). This model adopts the ratio of gas permeability to no-slip permeability can be
a linear combination of gas-transport mechanisms to predict the expressed as
overall flow rate. It was mainly based on empirical observation
(Graham 1833, 1846, 1849, 1863), and the theoretical explanation  
kg  1 0:4
 4Kn
was proposed later (Mason and Malinauskas 1983). The total ¼ 1 þ 1:358tan ð4Kn Þ  Kn 1 þ : . . . . ð8Þ
mass flux of a single-size conduit is calculated as kl 1 þ Kn

402 August 2012 SPE Reservoir Evaluation & Engineering


14 adsorbed layer of interest here is on the organic material. This
layer diminishes in thickness as the pressure decreases.
12 We analyze the effect of the adsorbed layer on the single-
throat conductance. The ratio of the throat conductance without
the adsorbed layer to that of the in-situ condition (gas adsorbed at
10 in-situ temperature and pressure) is evaluated. The conductance
Fraction (%)

without the adsorbed layer is computed by assuming a liquid


8 phase is flowing. The purpose of this calculation is simply to iso-
late the geometric effect of the adsorbed layer. We assume the
6
thickness of the adsorbed layer is 0.7 nm at 28 MPa (Ambrose
et al. 2010) and decreases linearly with reducing pressure. We fur-
ther suppose the cross sectional area of the throat is circular, con-
4 sidering the SEM images (Ambrose et al. 2010; Wang and Reed
2009). For the calculation of liquid (i.e. no-slip) flow, we assume
2 the conductance of the circular tube depends on the fourth power
of characteristic size, in agreement with Stokes flow of a Newto-
0
nian fluid (Sakhaee-Pour 2012). The influence of the adsorbed
0 10 20 30 40 50 layer is important for throat sizes smaller than 50 nm, which con-
d (nm) stitute a significant fraction of shale pores. For a typical shale
sample, the characteristic throat size of 6 nm corresponds to the
Fig. 1—Pore-size distribution of a Barnett shale sample largest fraction of conduits, as shown in Fig. 1.
obtained from mercury-intrusion capillary pressure. We build our pore-space model on the basis of the SEM
images of shale and the drainage experiment. Some images of
This indicates that the nonlinearity of permeability increases with Barnett shale available in the literature are shown in Fig. 2. The
Kn. For the sake of convenience, we develop a polynomial form high-resolution images indicate many throats exist in organic
for the permeability enhancement for 0.1 < Kn < 0.8, which lies materials. The affinity of gas molecules for organic materials
within the range of interest for shale-gas reservoirs. A second- means that these throats constitute the void space that is altered
order polynomial works well: by desorption. It has been argued that the fluid flow in organic
materials is mainly single gas phase (Wang and Reed 2009).
kg Therefore, we propose a network model for gas permeability that
¼ 0:8453 þ 5:4576Kn þ 0:1633Kn2 ; . . . . . . . . . . . . . ð9Þ
kl includes only voids within the organic matter.
where the coefficients are calculated using the nonlinear regres-
sion model (Mendenhall et al. 1989), and the coefficient of regres- Results
sion is found to be 0.99. Note that Eq. 9 is valid only for We analyze the conductance of a single-sized throat to study the
transition flow regime. Also, the gas permeability is no longer a effects of the adsorbed layer and slippage. Then, we build a net-
linear function of Kn. This reveals that the Klinkenberg correction work model to examine these effects when the connected throats
cannot be employed for higher-Kn flow regimes. exhibit a distribution of sizes. Although the network does not ex-
plicitly represent the arrangement of voids in a shale, it does yield
a first-order estimate of difference in the gas-flow behavior
Free Molecular Regime. For completeness, we include the free through the shale at different conditions—for example, high pres-
molecular regime, which is relevant for gas-phase transport at am- sure (early in the life of a well) vs. low pressure (after substantial
bient conditions. The mass-flow rate of the free molecular regime gas production) or field condition vs. laboratory condition. We
was modeled by Knudsen (1909). This model implements flux take into account the effects of the adsorbed layer and slippage on
because of a density gradient of the molecules as follows: the conductance of each throat in the network. Consequently, we
develop a characteristic plot, mapping the measurements taken at
JKn ¼ DKn rni ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð10Þ typical laboratory conditions to the in-situ conductance of each
throat.
where JKn is the mass flux of component i, DKn the Knudsen diffu-
sivity coefficient, and rni the density gradient. Unlike ordinary Analysis of a Single Cylindrical Conduit. Before considering
diffusion, only one component is considered to predict the flow the network, we investigate the effects of the adsorbed layer and
rate in this mechanism. This model requires Knudsen diffusivity, flow slippage on the conductance of a single throat. In this regard,
which is usually measured experimentally (Reinecke and Sleep we consider pore-size distribution (inferred from mercury poros-
2002). It was derived analytically only for a long tube with a con- imetry) and SEM of the Barnett shale sample shown in Figs. 1
stant circular cross-sectional area (Roy et al. 2003), as follows: and 2, respectively. We assume the void space is a network of cy-
rffiffiffiffiffiffiffiffiffi lindrical throats. Our analysis (Sakhaee-Pour 2012) shows qualita-
d 8RT d tively similar flow behavior, in terms of response to the adsorbed
DKn ¼ ¼ v; . . . . . . . . . . . . . . . . . . . . . . ð11Þ
3 pm 3 layer and slippage, if the network elements are slits rather than
cylinders.
where d is the tube diameter, m the molar mass, T the ambient We begin by analyzing the effect of the adsorbed layer on the
temperature, R the universal gas constant that is equal to 8.314 in liquid conductance of a single-sized throat. The goal is simply to
SI units, and v the average velocity of the molecules. quantify the geometric effect of reducing the throat size as the
thickness of the adsorbed layer increases. Thus, slip is ignored;
the calculations assume no slip on the adsorbed layer. The role of
Adsorbed Layer pressure in this analysis is merely to change the thickness of the
The effect of adsorbed gas is neglected in modeling flow through layer. The results are presented in terms of the ratio of the throat
conventional rock. This is reasonable because the occupied vol- conductance without the adsorbed layer to the throat conductance
ume is negligible compared to the total void space in conventional with the adsorbed layer (see Fig. 3). We observe that the adsorbed
rocks. However, the adsorbed volume of CH4 is crucial in shale layer is crucial in estimating the conductance (with a no-slip
because the throats are often smaller than 10 nm and much of the boundary condition), especially at larger pressures and in smaller
void space is in the organic material, for which CH4 has a large pore throats in the ranges of interest. The ratio of liquid conduc-
affinity. While molecular adsorption can occur on any surface, the tances is larger at smaller throat sizes because the layer thickness,

August 2012 SPE Reservoir Evaluation & Engineering 403


8 P=28 MPa
P=20 MPa
7
P=10 MPa
P=5 MPa
6

k(Kn=0)/kads
5

1
101
d (nm)

Fig. 3—Effect of adsorbed CH4 layer(s) on hydraulic conduct-


ance of a single cylindrical conduit. The ratio of “liquid” con-
ductance without adsorbed layer [k(Kn 5 0)] to “liquid”
conductance with adsorbed layer (kads) increases as pressure
increases and as conduit diameter decreases. Liquid conduct-
ance refers to single-phase laminar flow with no-slip boundary.
Thickness of adsorbed layer is proportional to pressure.

of gas to liquid conductances is shown in Fig. 4. The calculation


reveals that slippage plays an important role in smaller throat
sizes and at lower pressures. For instance, the gas conductance
of a 6-nm throat, which has the largest fraction of throat sizes in
Fig. 1, is larger than the liquid conductance by a factor of 1.5 at
P ¼ 20 MPa. This ratio increases nonlinearly with the decrease of
pressure, revealing that the slippage becomes more important at
late production.
Now, we consider the combination of the competing effects of
the adsorbed layer and slippage on the gas conductance. The
adsorbed layer reduces the conductance by reducing the cross-sec-
tional area. Slippage enhances the conductance by facilitating the
molecule movements at the throat surface. Both effects depend on
pressure. To analyze this competition, we calculate the ratio of
gas conductance with the adsorbed layer to liquid conductance
without the adsorbed layer. As in the previous analysis, the liquid
conductance is calculated with a no-slip boundary condition to
serve as a reference value. The gas conductance is obtained by

3 P=28 MPa
P=20 MPa
P=10 MPa
k[Kn(P,T=300 K)]/k(Kn=0)

P=5 MPa
2.5

Fig. 2—SEM of Barnett shale (Ambrose et al. 2010; Wang and 1.5
Reed 2009).

which is constant at a given pressure, occupies a greater fraction 1


101
of the cross section in smaller throats.
d (nm)
Next, we explore the importance of slippage with respect to
the gas conductance. To this aim, we calculate the ratio of gas to
Fig. 4—Effect of slip at pore walls on hydraulic conductance of
liquid conductances. The conductances are computed assuming a single cylindrical conduit [ratio of gas conductance with slip-
that no adsorbed layer exists. The only effect of pressure is thus to page (Kn evaluated at 300 K and at pressure as per legend) to
change Kn and, hence, the degree of slip. The temperature is fixed liquid conductance without slippage k(Kn 5 0)]. Both conduc-
at 300 K. As before, the liquid conductance is computed with a tances assume no adsorbed layer is present. The value of the
no-slip boundary condition to serve as a reference value. The ratio Kn changes with pressure.

404 August 2012 SPE Reservoir Evaluation & Engineering


3.5 P=28 MPa 12 P=28 MPa
P=20 MPa P=20 MPa

klab[Kn(P=5 MPa, T=300 K)]


kin-situ[Kn(P,T=360 K)]/k(Kn=0)
3 P=10 MPa 10 P=10 MPa

/kin-situ[Kn(P,T=360 K)]
P=5 MPa
2.5 8

2 6

1.5 4

1 2

0.5 0
101
0 d (nm)
101
d (nm) Fig. 6—The ratio of gas conductance klab without adsorbed
layer and with slippage (laboratory condition of 300 K and 5
MPa) to gas conductance kin-situ with adsorbed layer and slip-
Fig. 5—Combined effect of adsorbed layer and slip on hydraulic
page (field condition of 360 K and P as per legend) increases as
conductance of a single cylindrical conduit [ratio of gas con-
pressure increases and as cylindrical throat size decreases.
ductance kin-situ with adsorbed layer and slippage (in-situ con-
These curves are relevant to estimating field permeability from
dition of 360 K and P as per legend) to liquid conductance
laboratory measurements.
k(Kn 5 0) without adsorbed layer and without slippage]. This
behavior is relevant to evolution of permeability during
production. tory condition is notably higher than at the in-situ condition at
larger pressures. This is because of the lack of an adsorbed layer
and the enhancement of slippage at laboratory condition vs. in-situ
computing slip flow within the open cross section of the throat condition. Clearly, the laboratory measurements should be cor-
remaining after implementing the adsorbed layer at the specified rected for any flow modeling at the field condition. Considering the
pressure. The temperature is held constant at 360 K. Fig. 5 shows importance of this issue, we will analyze the laboratory measure-
qualitatively different trends, depending on pressure. At large ments using network modeling and propose a correction in the fol-
pressures, the ratio of gas conductance at in-situ conditions to the lowing paragraphs.
corresponding no-slip conductance increases as the pore size Liquid permeability is also of interest in shale, either for the
increases. At small pressures, the ratio decreases as the pore size problem of water production or for the challenge of “tight oil”
increases. production. To estimate liquid permeability at the field condition
This calculation helps us understand which effect governs the from gas permeability measured at the laboratory condition, we
flow behavior at different stages of production. The purpose of calculate the ratio of gas to liquid conductances of a single cylin-
the comparison between gas and liquid conductances is to analyze drical conduit. As in the preceding, the adsorbed layer at the labo-
the gas-conductance variation vs. a reference value. It compares ratory condition is negligible because of low pressure. Therefore,
the gas conductance (360 K, various P) that is representative of the gas conductance at the laboratory condition is computed from
reservoir condition to liquid conductance, the reference value the liquid conductance by implementing the slippage at the labo-
evaluated in the absence of an adsorbed layer and ignoring slip. ratory pressure and temperature (5 MPa, 300 K). For the liquid
The analysis shows the adsorbed layer dominates the flow behav- conductance at the in-situ condition, we calculate the throat area
ior at high pressure (P ¼ 28 MPa in Fig. 5) regardless of the throat not obstructed by the thickness of the adsorbed layer. The thick-
size. That is, the gas conductance is smaller than the reference ness of the adsorbed layer depends on pressure. We ignore the
conductance for all throat sizes because the adsorbed layer signifi- swelling effect for the in-situ condition, which may reduce the hy-
cantly reduces the cross section for gas flow. The influence is draulic conductance. This may occur because of the presence of
greater for smaller throats. However, the slippage has the primary water in the formation, which is not available at the same amount
effect at small pressures (P ¼ 10 MPa or P ¼ 5 MPa). That is, the at the laboratory condition. The ratio of gas to liquid conductan-
gas conductance is greater than the reference conductance for all ces at laboratory and field conditions is shown in Fig. 7. The
throats, with the effect being greater in narrower throats. The results reveal the gas permeability obtained at the laboratory con-
implication of Fig. 5 is that the adsorbed layer should be taken into dition is significantly larger than the liquid permeability at the in-
account for modeling gas transport during early production, while situ condition. The ratio of permeabilities estimated here is larger
the influence of slippage is dominant during later production. than in the preceding analysis (Fig. 6) because the liquid does not
Finally, for the single-tube study, we compare the gas conduc- slip at laboratory condition. In reality, the ratio of gas to liquid
tances at laboratory and in-situ conditions. This is useful for conductances may be even greater than in Fig. 7 because the hy-
estimating gas permeability in the field from a laboratory measure- draulic conductance of the throat decreases because of possible
ment. The adsorbed layer is presumed negligible at the laboratory swelling clays available in the formation.
condition because the pressure (5 MPa) is low; the curve for the
corresponding pressure in Fig. 3 shows that the correction is less Analysis of Network of Cylindrical Conduits of Distributed
than 25% for all throat sizes. The thickness of the adsorbed layer Sizes. We employ a regular square lattice network (Bennett and
varies at the in-situ condition, depending on the pressure. Note that Myers 1962) to model the gas flow through organic materials in
the slippage occurs in both conditions. We calculate the gas con- shale. We adopt the pore-size distribution from the drainage
ductance at the laboratory condition (T ¼ 300 K) from the liquid experiment shown in Fig. 1. We also assume the width of the net-
conductance, accounting for slippage at laboratory pressure in the work (transverse to the direction of flow) and the throat length are
relevant flow regime (Kn criterion evaluated at 300 K and 5 MPa) 17 and 50 nm, respectively. The model width is chosen so that the
by means of Eq. 2, Eq. 3, or Eq. 9. For the in-situ condition, we network model yields a desired porosity. Here, 17 nm is the value
compute the cross-sectional area open to flow, depending on the that results in the desired porosity of 10%. The small width means
thickness of the adsorbed layer. The gas conductance is computed that the network corresponds to a thin 2D slice through the or-
from the liquid conductance at the field pressure after implement- ganic material oriented along the axes of the circular holes. That
ing the slippage. Fig. 6 shows that the gas conductance at labora- is, we consider flow perpendicular to the plane of the images in

August 2012 SPE Reservoir Evaluation & Engineering 405


25 P=28 MPa 4.5
klab[Kn(P=5 MPa, T=300 K)]/kads P=20 MPa
4
P=10 MPa
20
P=5 MPa 3.5

kg, lab/kg, in-situ


15 3

2.5
10
2

5 1.5

1
0
1
10 0.5
d (nm) 5 10 15 20 25 30
P (MPa)
Fig. 7—For a single cylindrical tube, a laboratory measurement
of gas permeability greatly overestimates the permeability to
liquid in the reservoir, as shown by the ratio of gas conduct- Fig. 8—The ratio of gas permeability with slippage and without
ance klab with slippage and without adsorbed layer (laboratory adsorbed layer (laboratory condition of 300 K and 5 MPa) to gas
condition of 300 K and 5 MPa) to liquid conductance kads with- permeability with adsorbed layer and slippage (field condition
out slippage and with adsorbed layer (field condition of 360 K of 360 K and P from the x-axis) obtained from a network of cy-
and P as in legend). Pressure determines the thickness of the lindrical throats having the size distribution shown in Fig. 1.
adsorbed layer.

N2 than for CH4 because of a smaller Kn, resulting in a smaller


permeability. Thus, the gas permeability of the network model
Fig. 2. The objective is not to represent the actual connectivity of exposed to CH4 at a slightly larger pressure than 5 MPa at field
the voids, but to capture the influence of the adsorbed layers and condition is equal to the N2 permeability at the laboratory
slip in a connected collection of throat sizes. When the throat condition.
sizes are distributed randomly on the network, the gas permeabil- The ratio of laboratory to in-situ permeabilities decreases as
ity of the network is 93 nanodarcies at laboratory condition [i.e., the pressure decreases. This is because of the opposing effects of
without the adsorbed layer; the calculation accounts only for slip the adsorbed layer and slippage at different pressures. To provide
in each throat, using Eqs. 9 and 3 (high Knudsen flow models) to a quick tool for estimating the in-situ permeability from the labo-
relate flow in each throat to the pressure gradient along that ratory measurement, we fit a curve to the results in Fig. 8. The
throat]. The network permeability is a plausible overestimate of nonlinear regression model indicates the coefficient of regression
the measured value because the network represents only the or- is 0.99. Therefore, the in-situ gas permeability can be estimated
ganic portion of the shale. The organic matter accounts for 10% from the laboratory measurement using the following equation:
of the core volume, so prorating the network contribution accord-
ingly would yield an estimate of 9 nanodarcies, which is typical kg;lab
¼ 0:001P2 þ 0:0898P þ 0:783; . . . . . . . . . . . ð12Þ
for shale (Billiotte et al. 2008). kg;in-situ
Effect of Laboratory Conditions vs. Field Conditions on Gas
Permeability. First, we evaluate the gas permeability at in-situ where kg,in-situ is the gas permeability at the in-situ condition, kg,lab
conditions by considering the influence of the adsorbed layer on is the gas permeability at the laboratory condition (300 K, 5
flow through each conduit in the network. To compute the gas MPa), and P is the gas pressure in MPa.
conductance at the in-situ condition, we calculate the liquid con- Effects of Field Conditions on Gas Permeability During
ductance with no-slip boundary condition based on the cross-sec- Production. Finally, we investigate the effect of declining field
tional area, depending on the thickness of the adsorbed layer. The pressure on the gas permeability using the network model. In this
gas conductance of each throat is then computed from the liquid regard, we calculate the ratio of the field permeability at low pres-
conductance. sures (kg2,in-situ) to the gas permeability value at the start of pro-
The ratio of network permeabilities at laboratory conditions to duction (kg1,in-situ). The initial field pressure is presumed to be 28
those at in-situ conditions is presented in Fig. 8. The results show MPa. We take into account the effect of the adsorbed layer,
the measurements made at the laboratory condition would overes- depending on the pressure. In addition, we account for the effect
timate the permeability by a factor of four at P ¼ 28 MPa (i.e., at of pressure on slippage. The ratio of the gas permeabilities is pre-
the start of production). This factor is almost equal to the ratio of sented in Fig. 9. The network modeling suggests that the shale
laboratory to in-situ permeabilities for a single tube of 6 nm (see permeability at the start of production is significantly smaller than
Fig. 6), which has the largest fraction of pore-throat sizes as during late production. The adsorbed layer is the reason for this
deduced from the mercury-intrusion data shown in Fig 1. This effect. Furthermore, the dependency of the permeability on the
means that the throats, of which characteristic sizes constitute the pressure is nonlinear. Hence, the gas permeability of shale is
largest population, dominate the flow behavior. The overestima- highly dependent on pressure, unlike the conventional reservoir.
tion would be even larger if the prediction were based on labora- This has major implications in reservoir simulations and ultimate-
tory measurements at ambient pressure and temperature. This is recovery-estimation models.
because of enhancement in slippage occurring at low pressures. Validation Against Laboratory Data. Figs. 8 and 9 summa-
The ratio reaches unity at a slightly larger pressure than 5 MPa rize the main implications of the theory. Fig. 8 estimates gas
because the field temperature (360 K) is greater than the labora- permeability at the in-situ condition from the laboratory measure-
tory temperature (300 K) and because the gas for the laboratory ment, and Fig. 9 predicts the enhancement of gas permeability
measurement is N2 and not CH4. The difference in temperatures during production. Two sets of experiments are required to vali-
makes Kn slightly larger at the laboratory condition, but for a date these implications. The first set requires comparison of the
given pressure, temperature, and throat size, the properties of N2 measurements with N2 at the laboratory condition and CH4 at the
yield a smaller Kn than for CH4. Thus, smaller slippage occurs for reservoir condition. For the second set, we must analyze the

406 August 2012 SPE Reservoir Evaluation & Engineering


4.5 1.9

1.8 Laboratory data (Letham 2011)


4
1.7 Corrected throats model

klab(P, T=300 K)/k(Kn=0)


3.5
kg 2, in-situ /kg1, in-situ
1.6
3
1.5

2.5 1.4

1.3
2
1.2
1.5
1.1
1 1
5 10 15 20 25 30 0 1 2 3 4 5 6 7
P (MPa) P (MPa)

Fig. 9—The ratio of gas permeability at pressures below initial Fig. 10—The ratio of gas permeability without an adsorbed
reservoir pressure (kg2,in-situ) to gas permeability at initial pres- layer and with slippage [klab(P, T5300 K)] to the liquid perme-
sure (P 5 28MPa) (kg1,in-situ) increases as production continues ability without an adsorbed layer and without slippage
and pressure declines accordingly. The ratio is calculated from [k(Kn 5 0)] is obtained from a network of conduits and com-
a network of cylindrical throats having the size distribution pared to the laboratory measurements (Letham 2011). The
shown in Fig. 1. effective stress is constant, and hence, the pore-pressure
change does not affect pore-throat-size distribution. Gas per-
meability increases with lowering pressure because of slip-
page, unlike the liquid permeability, which is constant. The
dependency of rock resistance against gas flow on the pore pres- flowing gas is methane.
sure when the moving fluid is methane. The test with methane
should be run at reservoir temperature and pressures to represent
the in-situ conditions. Our model predicts that the effect of the mates the normalized gas permeability, the maximum difference
adsorbed layer becomes more important as the organic content of with the laboratory measurement being only 6%. This indicates
shale increases. The organic-rich region has more tendancy to that the slippage model provides a good estimate for the effect of
adsorb methane to its pore wall, which must be considered when pressure on gas permeability at moderate pressures.
choosing the core samples.
Data across the range of conditions of Figs. 8 and 9 for rocks
with different organic content are not currently available. How- Conclusions
ever, measurements of CH4 transport reported on a particular type Pore throats of shale are mostly narrower than 10 nm and are
of shale at pressures below 7 MPa (Letham 2011) let us partially inside organic material on which CH4 adsorbs. As a result, the gas
test the effect of slippage. The experimental data were measured permeability of these rocks is affected significantly by the
with methane for a sample with a permeability of 390 nanodarcies adsorbed gas and by the slip of flowing gas on the pore walls (or
at 5 MPa and 300 K. The effective stress was kept constant in the on layers of molecules adsorbed on the walls). The adsorbed layer
laboratory measurements (Letham 2011); that is, the confining does not play an important role in conventional rocks because the
stress was lowered to yield the same effective stress at a lower pore throats are much wider, nor is the gas volume desorbed from
pore pressure. The porosity of the laboratory sample was reported the organic material significant in conventional reservoirs. To bet-
to be 13%. The porosity of our original sample was 10%, which is ter understand gas-flow behavior through shale, we evaluated
slightly different from the laboratory sample. To compensate for these effects in individual conduits of a cylindrical cross section
these differences, we increased the pore-throat sizes of our origi- and in simple networks of such conduits. The effect of the
nal network model, of which permeability adsorbed layer was treated as purely geometric: The cross section
qwas
ffiffiffiffiffiffi 9 nanodarcies

at 5
open to flow was reduced by the thickness of the adsorbed layer,
390
MPa and 300 K, by a factor of 6.58 9 ¼ 6:58 . We also which was assumed to vary linearly with pressure. The effect of
increased the width of the network by a factor of 43.3 (¼ 6.582) to slip was accounted for by applying the model appropriate to the
keep the porosity of the network model unaffected by the change flow regime, according to the Knudsen number (Kn). The latter
in the pore-size distribution. Using the modified model, we com- depends on pressure, temperature, and conduit diameter. For the
pute gas permeabilities for a range of small pressures (between 1 slip flow regime, the first-order slip model was judged more suita-
and 6 MPa) and at 300 K. Since the effective stress was kept con- ble than the dusty gas model (DGM) for the shale-gas application.
stant during laboratory measurements, we do not change the At large pressures such as typical initial shale-gas-reservoir
modified pore-throat-size distribution with pore pressure. (In the pressures, the effect of the adsorbed layer dominates the effect of
field, effective stress increases with reservoir depletion, and the slip on gas-phase permeability. Slip dominates at smaller pres-
effect on pore-throat-size distribution, which is not included in sures typical of those after longer periods of production. Conse-
the model presented here, would diminish the slippage effect in quently, the reservoir matrix permeability is predicted to increase
Fig. 10.) Fig. 10 plots these gas permeabilities, normalized by the significantly throughout the life of a well, by a factor of 4.5, as
nominal liquid permeability at laboratory condition. Similar to ev- production continues and pressure declines. The models predict
ery calculation performed for the laboratory condition through that the typical conditions for laboratory measurements of perme-
this study, the gas flow here is with slippage and without the ability cause those values to overestimate field permeability by as
adsorbed layer, and the liquid flow is without slippage and with- much as a factor of four. The model results are captured in simple
out the adsorbed layer. However, unlike the laboratory condition analytical expressions that allow convenient estimation of these
we defined in which pressure was 5 MPa, we change the pressure effects.
to investigate its influence on gas permeability because of slip- For complete validation of the proposed model, laboratory
page. The liquid permeability for the laboratory data is 390 nano- measurements at elevated pressures (on the order of 28 MPa) and
darcies, estimated by extrapolating the measured gas permeability reservoir temperature made with CH4 are needed. Experimental
to large pressure. Fig. 10 shows that the model slightly underesti- data at such conditions, for which the adsorbed layer is expected

August 2012 SPE Reservoir Evaluation & Engineering 407


to have the dominant effect, are not available. Comparing model burgh, Pennsylvania, USA, 23–25 February. http://dx.doi.org/10.2118/
predictions with the laboratory data at lower pressures (<7 MPa) 131772-MS.
and constant effective stress permits evaluating the importance of Bennett, C.O. and Myers, J.E. 1962. Momentum, Heat, and Mass Transfer.
slippage. When normalized by the nominal absolute permeability New York: Series in Chemical Engineering, McGraw-Hill.
of the network (i.e., the value at which both slip and adsorbed Billiotte, J., Yang, D., and Su, K. 2008. Experimental study on gas permeabil-
layers are negligible), the predicted trend agrees well with the ity of mudstones. Physics and Chemistry of the Earth, Parts A/B/C 33
measurements. The maximum difference between the predicted (Supplement 1): S231–S236. http://dx.doi.org/10.1016/j.pce.2008.10.040.
normalized gas permeability and the measured value is 6%. This Cipolla, C.L., Lolon, E.P., Erdle, J.C. et al. 2010. Reservoir Modeling in
means that our adapted network model provides a reasonable ba- Shale-Gas Reservoirs. SPE Res Eval & Eng 13 (4): 638–653. SPE-
sis for understanding the effect of gas pressure on the matrix per- 125530-PA. http://dx.doi.org/10.2118/125530-PA.
meability of a shale-gas reservoir. Cui, X., Bustin, A.M.M., and Bustin, R.M. 2009. Measurements of gas
permeability and diffusivity of tight reservoir rocks: different
approaches and their applications. Geofluids 9 (3): 208–223. http://
Nomenclature
dx.doi.org/10.1111/j.1468-8123.2009.00244.x.
d ¼ tube diameter Graham, T. 1833. On the law of the diffusion of gases. J. Membr. Sci. 100
DKn ¼ Knudsen diffusivity coefficient (1): 17–21. http://dx.doi.org/10.1016/0376-7388(94)00228-Q.
JKn ¼ Knudsen flux Graham, T. 1846. On the Motion of Gases. Phil. Trans. R. Soc. Lond. A
Jtotal ¼ total flux 136 (1846): 573–631.
Jvisc ¼ viscous flux Graham, T. 1849. On the Motion of Gases. Part II. Phil. Trans. R. Soc.
k(Kn ¼ 0) ¼ no-slip conductance for a single cylin- Lond. A 139 (1849): 349–391.
drical tube with no adsorbed layer Graham, T. 1863. On the Molecular Mobility of Gases. Phil. Trans. R.
k[Kn(P,T ¼ 300 K)] ¼ gas conductance for a single cylindri- Soc. Lond. A 153 (1863): 385–405.
cal tube with slippage and without an Javadpour, F., Fisher, D., and Unsworth, M. 2007. Nanoscale Gas Flow in
adsorbed layer
Shale Gas Sediments. J Can Pet Technol 46 (10): 55–61. JCPT Paper
kads ¼ no-slip conductance for a single cylin-
No. 07-10-06. http://dx.doi.org/10.2118/07-10-06.
drical tube in the presence of an
Karniadakis, G., Bes kök, A., and Aluru, N. 2005. Microflows and Nano-
adsorbed layer
flows: Fundamentals and Simulation, Vol. 29. New York: Interdisci-
kg ¼ single-phase gas permeability
plinary Applied Mathematics, Springer. ISBN 0387221972.
kg,in-situ ¼ gas permeability of the network
Klinkenberg, L.J. 1941. The permeability of porous media to liquids and
model at the in-situ condition
gases. API Drilling & Production Practice (1941): 200–213.
kg,lab ¼ gas permeability of the network
Knudsen, M. 1909. Die Gesetze der Molekularströmung und der inneren
model at the laboratory condition
Reibungsströmung der Gase durch Röhren (The laws of molecular and
kg1,in-situ ¼ gas permeability of the network
viscous flow of gases through tubes). Ann. Phys. 333 (1): 75–130.
model at the start of production repre-
senting the in-situ condition http://dx.doi.org/10.1002/andp.19093330106.
kg2,in-situ ¼ gas permeability of the network Letham, E.A. 2011. Matrix permeability measurements of gas shales: gas
model at a different reservoir pressure slippage and adsorption as sources of systematic errors. BSc thesis,
representing the in-situ condition University of British Columbia, Vancouver, Canada.
kin-situ[Kn(P,T ¼ 360 K)] ¼ gas conductance for a single cylindri- Luffel, D.L., Hopkins, C.W., and Schettler Jr., P.D. 1993. Matrix Perme-
cal tube at the in-situ condition with ability Measurement of Gas Productive Shales. Paper SPE 26633 pre-
slippage and an adsorbed layer sented at the SPE Annual Technical Conference and Exhibition,
kl ¼ permeability of a conduit to liquid Houston, 3–6 October. http://dx.doi.org/10.2118/26633-MS.
with no-slip boundary condition Mallon, A.J. and Swarbrick, R.E. 2008. How should permeability be meas-
klab[Kn(5 MPa,300 K)] ¼ gas conductance for a single cylindri- ured in fine-grained lithologies? Evidence from the chalk. Geofluids 8
cal tube at the laboratory condition (1): 35–45. http://dx.doi.org/10.1111/j.1468-8123.2007.00203.x.
with slippage and without an adsorbed Mason, E.A. and Malinauskas, A.P. 1983. Gas Transport in Porous
layer Media: The Dusty-Gas Model. Amsterdam, The Netherlands: Chemi-
Kn ¼ Knudsen number cal Engineering Monographs, Elsevier Science.
m ¼ molar mass Mendenhall, W., Reinmuth, J.E., and Beaver, R. 1989. Statistics for Man-
M_ ¼ mass flux agement and Economics. Boston, Massachusetts: PWS-Kent Publish-
p ¼ gas pressure ing Company.
P ¼ gas pressure in MPa Philip, Z.G., Jennings Jr., J.W., Olson, J.E. et al. 2005. Modeling Coupled
R ¼ universal gas constant Fracture-Matrix Fluid Flow in Geomechanically Simulated Fracture
R ¼ specific gas constant Networks. SPE Res Eval & Eng 8 (4): 300–309. SPE-77340-PA.
T ¼ temperature http://dx.doi.org/10.2118/77340-PA.
a1 ¼ permeability enhancement Polczer, S. 2009. Shale expected to supply half of North America’s gas.
k ¼ mean free path of gas molecule Calgary Herald, 9 April 2009.
K ¼ characteristic length of the channel Reinecke, S.A. and Sleep, B.E. 2002. Knudsen diffusion, gas permeability,
l ¼ gas viscosity and water content in an unconsolidated porous medium. Water Resour.
q ¼ gas density Res. 38 (12): 1280. http://dx.doi.org/10.1029/2002wr001278.
Roy, S., Raju, R., Chuang, H.F. et al. 2003. Modeling gas flow through
microchannels and nanopores. J. Appl. Phys. 93 (8): 4870–4879.
http://dx.doi.org/10.1063/1.1559936.
Acknowledgments
Sakhaee-Pour, A. 2012. Gas flow through shale. PhD dissertation (in pro-
The authors acknowledge financial contributions from Exxon- gress), The University of Texas at Austin, Austin, Texas.
Mobil and Jackson School of Geosciences at The University of Urbina, I. 2011. Insiders Sound an Alarm Amid a Natural Gas Rush. N.Y.
Texas at Austin. Times, 25 June 2011, http://www.nytimes.com/2011/06/26/us/26gas.
html?_r¼1&pagewanted¼all.
References Wang, F.P. and Reed, R.M. 2009. Pore Networks and Fluid Flow in Gas
Ambrose, R.J., Hartman, R.C., Diaz-Campos, M. et al. 2010. New Pore- Shales. Paper SPE 124253 presented at the SPE Annual Technical
scale Considerations in Shale Gas in Place Calculations. Paper SPE Conference and Exhibition, New Orleans, 4–7 October. http://
131772 presented at the SPE Unconventional Gas Conference, Pitts- dx.doi.org/10.2118/124253-MS.

408 August 2012 SPE Reservoir Evaluation & Engineering


Appendix A ð ð1    
2  rv
First-Order Slip Model. The velocity profile in the slip flow re- Q ¼ us ðrÞ2prdr ¼ a 1  x2 þ 2 Kn xdx
gime for a cylindrical conduit is obtained by adding a correction rv
0
term to the solution of the Navier-Stokes equation after applying
the Newtonian-fluid assumption with no-slip boundary condition ð1      
Q 2  rv 1 2  rv
(Roy et al. 2003): Q¼ 1  x2 þ 2 Kn xdx ¼ þ Kn ;
a rv 4 rv
0
us ¼ uno-slip þ ucorrection ; . . . . . . . . . . . . . . . . . . . . . ðA-1Þ
                   ðA-8Þ
where us is the velocity profile in the slip flow regime, uno-slip is where Q is the flow rate, a is a characteristic flow rate, and Q is
the no-slip velocity of the Newtonian fluid, and ucorrection is the the nondimensional flow rate. The first term on the right-hand
correction term to implement the effect of slippage. The correc- side of Eq. A-8 represents the no-slip flow rate, and the second
tion term is constant, independent of location. The slip velocity term is the enhancement because of slippage. Therefore, the effect
profile can also be expressed in a nondimensional form as of slip on permeability can be readily measured by dividing
QðKn Þ by QðKn ¼ 0Þ :
Us ¼ Uno-slip þ Ucorrection ; . . . . . . . . . . . . . . . . . . . . ðA-2Þ

1 2rv  
where the nondimensional no-slip velocity is defined as follows: kg 4 þ rv Kn 2  rv
¼ 1
¼ 1 þ 4 Kn : . . . . . . . . . ðA-9Þ
r
2 kl 4
rv
Uno-slip ¼ 1  ¼ 1  x2 ; . . . . . . . . . . . . . . . . ðA-3Þ
R Here kg is the permeability of the conduit to gas phase, and kl is
where R is the radius of the tube, r is the distance from the center the permeability of the same conduit when the no-slip boundary
of the tube, and x is the normalized radial distance from the center condition applies. The latter condition usually holds when a liquid
of the tube, which is equal to unity at the wall. To include the fills the conduit, hence the choice of subscript l. After implement-
effect of molecule/wall collisions, the first-order slip boundary ing rv ¼ 0.9, we obtain the enhancement of apparent permeability,
condition is imposed as (Klinkenberg 1941) called a1 in the text (refer to Eq. 3), to be equal to 5.

2  rv @Us DGM Expression in Terms of Kn. The apparent permeability of
Us j w ¼ Kn ; . . . . . . . . . . . . . . . . . . . . ðA-4Þ
rv @n w the conduit based on the DGM is calculated as
 
where n is the outward (unit) vector, which is normal to the tube kg DK l b
¼ 1þ n ¼ 1 þ : . . . . . . . . . . . . . . . . . ðA-10Þ
wall, and rv is the tangential momentum-accommodation coeffi- kl kl p p
cient. The tangential momentum accommodation is calculated by
measuring the gas-flow rate at the slip flow regime and is close to The second term on the right-hand side of Eq. A-10 can be simpli-
0.9 (Roy et al. 2003). fied after substituting the Knudsen diffusivity coefficient in terms
The first-order slip boundary condition is applied to derive the of a pertinent parameter, using Eq. 11. We also use the Hagen-
nondimensional velocity profile (see Eq. A-2) (Klinkenberg Poiseuille model for the permeability term, as follows:
1941): rffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffi
  d 8RT 8RT
l
@Us @Uno-slip @Ucorrection @Uno-slip b DKn l 3 pm  l 32 1 pm
¼ þ ¼ ¼ ¼ ¼
@n w @n @n
w @n w p kl p d2 3 d p
p
¼ ð2xÞjx¼1 ¼ 2:                  ðA-5Þ 32
rffiffiffiffiffiffiffiffiffi
2RT
The minus sign is because of the direction of the normal vector. l
64 1 pm
The correction term is constant, and, thus, its gradient vanishes. ¼ :                    ðA-11Þ
Therefore, Eq. A-4 simplifies to (Klinkenberg 1941) 3 d p
 
2  rv @Us 2  rv Here, we implement the ideal-gas assumption p ¼ q mR T ¼ qRT
K ¼ 2 Kn : . . . . . . . . . . . . . . ðA-6Þ to express the pressure in terms of density, which yields the
rv @n w
n
rv results in terms of Kn:
Thus, we can express the nondimensional velocity profile based rffiffiffiffiffiffiffiffiffi
2RT rffiffiffiffiffiffiffiffiffi
on Eqs. A-2 and A-3 as follows (Klinkenberg 1941): l
b 64 1 pm 64 l 2m
  ¼ ¼
2  rv p 3 d R 3 qd pRT
U s ¼ 1  x2 þ 2 Kn : . . . . . . . . . . . . . . . . ðA-7Þ q T
rv r mffiffiffiffiffiffiffiffiffi
128 l p
To cast this profile in terms of permeability, we obtain the flow ¼ ¼ 13:58Kn :              ðA-12Þ
3p qd 2RT
rate as

August 2012 SPE Reservoir Evaluation & Engineering 409


What Is Stimulated Reservoir Volume?
M.J. Mayerhofer, Pinnacle; E.P. Lolon, CarboCeramics; N.R. Warpinski, Pinnacle; C.L. Cipolla, CarboCeramics;
D. Walser, Pinnacle; and C.M. Rightmire, Forrest A. Garb and Associates

Summary reservoirs because they maximize fracture-surface contact area


Ultralow-permeability shale reservoirs require a large fracture with the shale through both size and fracture density (spacing).
network to maximize well performance. Microseismic fracture Chances for creating large tensile-fracture networks are increased
mapping has shown that large fracture networks can be gener- by pre-existing healed or open natural fractures and favorable
ated in many shale reservoirs. In conventional reservoirs and tight stress-field conditions such as a small difference in principal hori-
gas sands, single-plane-fracture half-length and conductivity are zontal stresses. Fig. 3 shows an example of a complex-fracture
the key drivers for stimulation performance. In shale reservoirs, network (in a vertical well) measured with microseismic mapping.
where complex network structures in multiple planes are created, The figure illustrates the development of a large-scale network with
the concepts of single-fracture half-length and conductivity are two distinct, orthogonal fracture orientations. This paper aims to
insufficient to describe stimulation performance. This is the rea- expand on the papers cited by discussing in more detail the concept
son for the concept of using stimulated reservoir volume (SRV) of SRV and its relationship with shale-well performance.
as a correlation parameter for well performance. The size of the
created fracture network can be approximated as the 3D volume Estimating SRV
(stimulated reservoir volume) of the microseismic-event cloud. It has been well documented by Albright and Pearson (1982),
This paper briefly illustrates how the SRV can be estimated from Warpinski et al. (2004), and Rutledge and Phillips (2003) that
microseismic-mapping data and is then related to total injected- microseismic events are created mainly as a result of shear slip-
fluid volume and well performance. While the effectively produc- pages around the hydraulic fractures. The mechanisms include
ing network could be smaller by some proportion, it is assumed shear slippages induced by altered stresses near the tip of the
that the created and the effective network are directly related. fractures, and shear slippages related to leakoff-induced pore-
However, SRV is not the only driver of well performance. Fracture pressure changes. In conventional reservoirs with higher reservoir
spacing and conductivity within a given SRV are just as important, permeabilities and/or oil/water reservoirs, the microseismic-event
and this paper illustrates how both SRV and fracture spacing for a cloud can be fairly wide but may not be related to the generation of
given conductivity can affect production acceleration and ultimate complex-fracture networks. In this case, most of the events may be
recovery. The effect of fracture conductivity is discussed separately related to pore-pressure increase as a result of rapidly moving pres-
in a series of companion papers. Simulated-production data are sure transients in high-permeability formations and/or reservoirs
then compared with actual field results to demonstrate variability with fairly incompressible fluids. In supertight shale reservoirs,
in well performance and how this concept can be used to improve diffusivity-related pore-pressure changes cannot move very far
completion design, well spacing, and placement strategies. from the actual fracture planes, unless natural fractures in other
directions are opened and hydraulically enhanced as a network
Introduction structure, thus serving as a conduit for fluid movement. This means
Fisher et al. (2002), Maxwell et al. (2002), and Fisher et al. (2004) that a large event cloud structure must be approximately equivalent
were the first papers to discuss the creation of large fracture networks to the actual fracture-network size. Thus, the microseismic-event
in the Barnett shale and show initial relationships between treatment cloud structure observed by microseismic fracture mapping pro-
size, network size and shape, and production response. Microseis- vides a means to estimate the SRV in very tight reservoirs.
mic-fracture-mapping results indicated that the fracture-network size Fig. 4 shows an automated method using discrete bins to esti-
was related to the stimulation-treatment volume. Fig. 1 shows the mate stimulated reservoir area (SRA) from microseismic-mapping
relationship between treatment volume and fracture-network size for data in a horizontal well. Constant-width bins (i.e., 100-ft wide) are
five vertical Barnett wells, showing that large treatment sizes resulted drawn in the principal fracture direction from the wellbore to the
in larger fracture networks. It was observed that as fracture-network farthest event in the specific bin on both sides of the wellbore. The
size and complexity increase, the volume of reservoir stimulated individual-bin areas are then summed up to approximate the total
also increases. Fisher et al. (2004) detailed microseismic-fracture- SRA. The calculation of SRV (a 3D structure) requires an estimate
mapping results for horizontal wells in the Barnett shale. This work of the stimulated-fracture-network height in each discrete bin within
illustrated that production is related directly to the reservoir volume the contacted shale section, in addition to knowing the approximate
stimulated during the fracture treatments. In vertical wells, larger SRA. This calculation is also performed within the selected bins
treatments are the primary way to increase fracture-network size and and is performed by calculating the network height as the differ-
complexity. Horizontal-well geometry provides other optimization ence between the shallowest and deepest event within the specific
opportunities. Longer laterals and more stimulation stages can also bin and top and bottom of the shale section (Fig. 5). While this
be used to increase fracture-network size and SRV. Mayerhofer et al. method is not an analytically exact calculation, it does provide a
(2006) performed numerical reservoir simulations to understand fast automated method to approximate a very complex 3D structure,
the impact of fracture-network properties such as SRV on well while honoring the contact with the actual shale section. The SRV
performance. That paper also showed that well performance can be in this paper is specified in millions of ft3 or in acre-ft.
related to very long effective fractures, forming a network inside a An important aspect of the SRV measurement is the proper
very tight shale matrix of 100 nanodarcies or less. setup of observation well or multiple observation wells to guar-
Fig. 2 illustrates the various types of fracture growth, rang- antee that the entire SRV can be observed. The proper design
ing from simple fractures to very complex fracture networks. of a microseismic-mapping setup takes into account maximum
Complex-fracture networks are desirable in “supertight” shale observation distance to ensure that the entire SRV can be imaged.
Microseismic-moment magnitude vs. distance plots (Zimmer et al.
2007) can be used to ascertain if all events were within observable
Copyright © 2010 Society of Petroleum Engineers
range or if the SRV could in fact be larger than imaged. Other
issues that affect correlations between SRV and well performance
This paper (SPE 119890) was accepted for presentation at the SPE Shale Gas Production include associated formation-water production and condensate
Conference, Fort Worth, Texas, 16–18 November 2008, and revised for publication. Original
manuscript received for review 10 September 2008. Paper peer approved 27 March 2009. yield, which becomes relevant in less-mature shales close to the

February 2010 SPE Production & Operations 89


45000

40000

35000

Fracture Network Length (ft)


30000

25000

20000

15000

10000

5000

0
0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000
Fluid Volume (bbl)

Fig. 1—Relationship of total fracture-network length as a function of total job fluid volume pumped (Fisher et al. 2004).

oil window. Although general guidelines are applicable for SRV dimensions are given by the total fracture-network length (2xf) and
measurements, it is important to evaluate any given SRV measure- width (xn). However, the key property for well production is really
ment in conjunction with the particular reservoir setting. the total sum of all fracture-network segments (linear feet) within
the SRA, which is a strong function of fracture spacing (density).
SRV and Fracture Spacing in Shales Eq. 1 shows the calculation of total fracture length, L ftotal , for
It is important to note that the SRV is just the reservoir volume the entire SRV as a function of fracture-network half-length, xf ,
affected by the stimulation. It does not provide any details of the width, xn, and fracture spacing, Δxs:
effectively producing fracture structure or spacing. Maxwell et al.
(2006) introduced a concept that could eventually be used to charac- 4 x f xn
L ftotal = + 2 x f + x n . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
terize fracture density. In this approach, “additional seismic-signal ⌬x s
characteristics allow investigation of the source of the mechanical
deformation resulting in the microseisms. In particular, the seismic 1500
moment, a robust measure of the strength of an earthquake or All microseisms
microearthquake, can be used to quantify the seismic deformation.”
Besides this potential geophysical approach, reservoir modeling
1000 ~2,700 ft
also provides an avenue to evaluate the effectively producing net-
work better. The details of this approach will be described in the
following section. As an introduction, Fig. 6 shows an identical
SRV with different fracture spacings (densities) and the effect of 500
fracture spacing on the gas-recovery factor. The SRV in this graph is
approximately 2,000 × 106 ft3 (for h = 300 ft). In contrast to conven-
tional single-fracture modeling using fracture half-length, the SRA Frac
well
Northing (ft)

Simple fracture
Fracture Complex
Complexfracture
Fracture –500 ~1,200 ft

Possible
–1000
aligned
features
–1500

Observation well
–2000
–1000 –500 0 500 1000 1500
Complex fracture
Fracture Complex fracture
Fracture Easting (ft)
With fissure
with Fissureopening
Opening Network
network
Fig. 3—Microseismic fracture mapping shows complex-net-
Fig. 2—Types of fracture growth (Warpinski et al. 2008). work growth in shales (Warpinski et al. 2008).

90 February 2010 SPE Production & Operations


Fig. 4—Estimating SRA from microseismic-mapping data.

Fig. 7 shows a plot of Eq. 1 on a log-log plot for the given SRV
in Fig. 6. The resulting curve forms an approximate straight line
on a log-log plot. As an example, reducing fracture spacing from
Free spacing: 200 ft Free spacing: 100 ft
300 to 50 ft would result in a more than five-fold increase in total
fracture length (from 48,600 to 264,600 ft) and would accelerate
the 3-year cumulative gas recovery by about the same multiplier.
This illustrates the importance of viewing SRV in context with the

Xf
Treatment horizontal well
Xn After 1 year production
Gas Recovery Factor (%)

Events outside of shale section


Time (day)

Fig. 6—Effect of fracture spacing (⌬xs) on well performance


and recovery factor for SRV of approximately 2,000 × 106 ft3
Fig. 5—Conceptual sketch of SRV-height calculation. (Mayerhofer et al. 2008; Warpinski et al. 2008).

February 2010 SPE Production & Operations 91


1.E+06

Total Fracture Length (ft)

1.E+05

1.E+04
10 100 1000

Fracture Spacing (ft)

Fig. 7—Fracture spacing vs. total fracture length based on Eq. 1 for SRV of 2,000 × 106 ft3.

potential fracture spacing. A dual-porosity approach is not used at sizes (fluid volumes). While the data show some scatter, it clearly
this point because the fracture spacing is still relatively sparse and indicates a trend that large SRV’s will result in better well perfor-
the numerical approach better illustrates the linear flow patterns in mance. A similar trend is still valid when plotting SRV vs. longer
a low-permeability shale system. term 3-year cumulative gas production (Fig. 9). The scatter in the
data is to be expected as a result of variations in fracture spacing
SRV and Well Performance within the SRV and potential differences in effective network size
Previous publications have briefly discussed the relationship and conductivity. While the presented trends in Figs. 8 and 9 show
between actual SRV and well performance. Here we will expand a general relationship between SRV generated from microseismic
on the subject with more details and comparisons with longer-term mapping and production data, they do not provide any deterministic
production data from the Barnett shale. Fig. 8 shows a plot of quantification of how the actual effective network (portion of the
SRV measured from microseismic fracture mapping vs. 6-month network contributing to gas production) is structured to provide the
cumulative gas production from a group of horizontal wells in one specific gas production. The quantification of the effectively pro-
specific county in the Barnett shale. The plot shows that SRVs in ducing fracture network can be performed only with flow simula-
this county range from as low as 120 × 106 ft3 (2,755 acre-ft) to tions, such as those provided by a numerical reservoir simulator.
approximately 1,900 × 106 ft3 (43,618 acre-ft). Small SRVs were Previous work has provided extensive numerical reservoir mod-
generated partly as a result of shorter laterals and smaller treatment eling (Mayerhofer et al. 2006, Warpinski et al. 2008, Cipolla et al.

1000

900

800
6-month Cum Gas (MMscf)

700

600

500

400

300

200

100

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
SRV (Mio ft3)

Fig. 8—SRV trend vs. 6-month cumulative horizontal-well production for one Barnett-shale county.

92 February 2010 SPE Production & Operations


2500

2000

3-year Cum Prod (MMscf)


1500

1000

500

0
0 400 800 1200 1600 2000

SRV (Mio ft3)

Fig. 9—SRV trend vs. 3-year cumulative horizontal-well production for one Barnett-shale county.

2008) to show how specific fracture-network properties can affect fracture length within a microseismically mapped SRV of 219 ×
well performance. The work demonstrated that numerical model- 106 ft3 (gray dots in graph). Fracture spacing was modeled to be
ing using a black-oil reservoir simulator can history match well approximately 180 ft in the principal fracture direction, with a uni-
performance adequately. The numerical modeling describes the form fracture conductivity of 4 md-ft and shale-matrix permeability
fracture network as a discrete set of high-permeability fractures in of 100 nanodarcies (0.0001 md). The fracture network is larger on
two orthogonal directions. One direction represents the principal the southwest side of the wellbore, with only a few microseismic
hydraulic-fracture direction as a result of newly created fractures events to the northeast.
or the opening of natural fractures. The orthogonal set represents Fig. 11 shows the simulation setup for a 3,000-ft-long horizon-
the opening of conjugate natural-fracture sets. Besides geochemical tal well, from which an SRV of 5,000 × 106 ft3 with equidistant
shale properties and pore pressure, the key fracture-network proper- fracture spacing of 400 ft was generated. The plot clearly shows
ties that affect well performance include SRV, shale-matrix perme- that the ultimate drainage area is constrained to the created SRV in
ability, fracture spacing, and fracture conductivity. It was discussed supertight shale reservoirs. Using this approach, a series of simula-
in these papers that once the value for shale permeability is fixed, tions was generated with different SRV sizes and fracture spacings
the solutions for total cumulative fracture length (the sum of all to evaluate the effect of these properties on well performance. The
fracture segments in the network) and conductivity become fairly SRV was varied from approximately 800 × 106 ft3 to 6,000 × 106
unique. Fig. 10 shows the results from a vertical-well history match ft3 and equidistant fracture spacing from 50 to 800 ft.
(Mayerhofer et al. 2006), where the network was approximated with To cross check the generic simulations with actual well-production
an effectively producing network having almost 8,000 ft of total data and a more-detailed production history match, the vertical-well

Fig. 10—Reservoir-simulation results for vertical well (Mayerhofer et al. 2008).

February 2010 SPE Production & Operations 93


After 1 year After 15 years

k (matrix) = 0.0001 md

SRV = 5000 x 106 ft^3

2,450 ft

3,000 ft
Treatment Horizontal Well Treatment Horizontal Well

Fig. 11—Simulation of horizontal-well fracture network (SRV = 5,000 × 106 ft3, frac spacing = 400 ft, lateral length approximately
3,000 ft).

example in Fig. 10 was plotted against the more-generic simulation from 400 to 100 ft would result in a 3.5-fold increase in production.
results. Although the well configuration is different and flowing The larger the SRV, the more significant the fold of increase in
pressures are not exactly the same, the correlation between SRV, production with smaller fracture spacing (because of much longer
well performance, and fracture spacing is approximately consistent, total fracture length with larger SRV). On the other hand, generat-
showing that the 6-month cumulative production of this vertical ing a larger SRV with large 800-ft fracture spacing will not provide
well would fall on the 200-ft fracture-spacing curve (Fig. 12). increased production, which can be explained by the low fracture
Fig. 13 shows a plot of the same generic simulations with actual conductivity (and associated increasing pressure drop with distance
horizontal-well Barnett-shale data throughout the producing areas from the well) in a sparsely fractured network.
in north Texas. The figure shows a wide scatter of well results, After comparing short-term production, the discussion will now
with cumulative gas production ranging from less than 100 to 500 focus on longer-term production (3 to 15 years). Fig. 14 shows a
MMscf for SRVs ranging from 100 × 106 ft3 to 3,000 × 106 ft3. The plot of the SRV from the same generic horizontal-well simulations
data indicate that most of the wells plot at a fracture spacing of 200 vs. 3-year cumulative gas production, fracture spacing, and actual
ft or larger, with many wells at fracture spacings larger than 800 ft, horizontal-well Barnett-shale data. The data indicate a slight shift
indicating a sparse fracture network (smaller total fracture length) of the real data to smaller fracture spacings, which may indicate
within the measured SRV (Fig. 7) and therefore a less effective improved cleanup for the longer-term production. The actual cumu-
fracture network. The plot also shows the significant acceleration lative production ranges from as low as 100 to 2,000 MMscf. The
component of smaller fracture spacings on short-term production. simulations also show that the folds of increase in production as a
For example, at an SRV of 1,000 × 106 ft3, reducing fracture spacing function of smaller fracture spacing become less for longer-term

Vertical Well from Mayerhofer et al. 2006 falls on 200 ft frac spacing curve

Network spacing=400 ft Network spacing=200 ft Network spacing=100 ft


Network spacing=50 ft Vertical well Network spacing=800 ft

Fig. 12—Simulated SRV and actual 6-month vertical well production response (Mayerhofer et al. 2008).

94 February 2010 SPE Production & Operations


Network spacing=400 ft Network spacing=200 ft Network spacing=100 ft
Network spacing=50 ft Vertical well Network spacing=800 ft

Fig. 13—Simulated SRV and simulated 6-month cumulative production vs. actual Barnett-shale data (entire north Texas area).

recovery, as previously illustrated in Fig. 6 (early-time accelera- fracture spacing. For example, increasing SRV from 1,500 × 106
tion). The plot also clearly shows that significant improvements in ft3 to 3,000 × 106 ft3 with 800-ft fracture spacing will create no
well performance by generating larger SRVs is more significant at meaningful increase in gas production.
low to moderate SRVs (less than 1,000 × 106 ft3). For larger SRVs, Fig. 15 shows the results for 15-year cumulative produc-
the benefits can be achieved only with a simultaneous decrease of tion. In this case, no actual production data is available, because

3-Year Cumulative Gas vs. SRV


2000

1600
3-Year Cumulative Gas (MMscf)

1200

800

400

0
0 500 1000 1500 2000 2500 3000
SRV (Million cuft)
Network spacing=800 ft Network spacing=400 ft Network spacing=200 ft
Network spacing=100 ft Network spacing=50 ft Actual data

Fig. 14—Simulated SRV and simulated 3-year cumulative production vs. actual Barnett-shale data (entire north Texas area).

February 2010 SPE Production & Operations 95


15-Year Cumulative Gas vs. SRV
16000

Inputs:
Lateral length, ft: 3000
14000
Reservoir permeability, md: 0.0001 Largest observed SRV per well
Pore pressure, psi 3000
Net pay thickness, ft: 300
15-Year Cumulative Gas (MMscf)

12000 Frac conductivity, md-ft: 4


Minimum BHP, psi: 1000

10000

8000

6000

4000

2000

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000
SRV (Million cuft)
Network spacing=800 ft Network spacing=400 ft Network spacing=200 ft
Network spacing=100 ft Network spacing=50 ft

Fig. 15—Simulated SRV vs. 15-year cumulative production.

horizontal-well development started in 2002 in the Barnett shale. Applying SRV and Network Azimuth to
The plot shows that for longer-term production, the impact of SRV Well-Placement and -Spacing Strategies
becomes more significant even at larger fracture spacing because Figs. 16 and 17 show conceptual diagrams illustrating how SRV
SRV provides the ultimate drainage area and diffusion has had and fracture-network azimuth can be used for well-placement and
time to access the entire network. In this case, increasing SRV -spacing considerations, respectively. Fig. 16 shows the importance
from 1,500 × 106 ft3 to 3,000 x106 ft3 with 800-ft fracture spacing of lateral orientation with respect to fracture-network azimuth and
will result in a 50% cumulative-production increase from 2,000 to SRV. The graph assumes that the lateral will be drilled in a north-
3,000 MMscf (compared to no meaningful increase after 3 years west/southwest orientation. If fracture direction follows the same
of production). southwest-to-northwest trend, the scenario on the left side of Fig. 16

Fig. 16—Importance of SRV and network azimuth for lateral orientation and well placement.

96 February 2010 SPE Production & Operations


SRV 3
Adjacent SRV = Optimizing areal SRV 3 Overlapping SRV = Higher recovery factor
coverage but not recovery factor? (smaller frac spacing)

SRV 2
SRV 2 Well 3

Well 3
SRV 1
SRV 1 Well 2
Well 2

Well 1
Well 1

Fig. 17—Horizontal-well-placement strategy considerations in conjunction with SRV and fracture spacing in shale reservoirs.

will promote longitudinal fractures with a very small SRV. If wells critical to understand how they will impact well results, comple-
are to be drilled in this orientation, more wells with closer spacing tions and well-development strategies.
will be required to drain a given section of reservoir. The fracture 3. Engineering measures to increase SRV and fracture spacing
orientation provided in the center of Fig. 16 provides an oblique ori- include lateral length and orientation, treatment sizes, number
entation of the fracture direction with respect to the lateral and also of stages, perforation clusters, and diversion techniques and/or
results in less-efficient reservoir coverage, but SRV is now larger openhole packer completion systems. More perforation clusters
than in the longitudinal case. The best scenario with largest SRV and stages in cemented-and-cased completions increase the
is illustrated on the right side of Fig. 16, with transverse network likelihood of dense fracturing.
orientation providing the largest SRV per lateral. 4. The key completion strategy is to balance the creation of a large
Keeping these considerations in mind, one could postulate that SRV with maximum possible fracture density. Large SRVs with
well spacing (distance between laterals) should roughly coincide small fracture spacing provide maximum well performance
with the extent of the SRV perpendicular to the wellbore, as illus- and gas-recovery factors. However, the practical and physical
trated in Fig. 17 on the left-hand side (SRV from each lateral is limitations of achieving this goal will require an economic
adjacent to the offset well). This approach may appear to be the optimization process.
optimum configuration. While the areal coverage of SRV is defi-
nitely optimized with this approach, gas-recovery factors within Nomenclature
the SRV can be optimum only with sufficiently close fracture spac- hf = fracture height, L
ing, as illustrated in Fig. 6. If fracture spacing cannot be optimized xf = hydraulic-fracture wing or half-length, L
with large SRVs, it may be more beneficial to drill laterals at closer
xn = hydraulic-fracture network width (from microseismic-
spacing, with the chance of overlapping SRVs (right-hand side in
event pattern), L
Fig. 17). This is the reason for the so-called “simul-fracs” and “zip-
per fracs” in shale reservoirs, where multiple parallel wellbores are ⌬xs = orthogonal fracture spacing, L
either stimulated simultaneously or in short alternating sequence L ftotal = total composite fracture length
from one lateral to the next. The hypothesis in these strategies
is to maximize fracture density between laterals as a result of Acknowledgments
stress diversion from simultaneously growing opposing fractures The authors would like to thank Pinnacle (a Halliburton Service)
(simul-fracs) and/or creating new branches as a result of increased and Carbo Ceramics for supporting the publication of this work.
stresses around newly opened fractures. Real data from Barnett- Some of this work is also based on previous publications of Devon
shale completions may suggest that this approach has merits, but Energy Corporation data, and we thank them for their support of
no specific study has been published on this topic. The challenge those efforts.
lies in understanding the practical and physical limitations of
what is possible in terms of network size and fracture spacing at References
reasonable completion costs (i.e., lateral length, number of stages Albright, J.N. and Pearson, C.F. 1982. Acoustic Emissions as a Tool for
and perforations, and treatment size). Hydraulic Fracture Location: Experience at the Fenton Hill Hot Dry Rock
Site. SPE J. 22 (4): 523–530. SPE-9509-PA. doi: 10.2118/9509-PA.
Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J., Lolon, E.P., and Vincent,
Conclusions M.C. 2008. The Relationship between Fracture Complexity, Reservoir
1. Low-permeability shale reservoirs require a large fracture Properties, and Fracture Treatment Design. Paper SPE 115769 pre-
network with small fracture spacing and adequate conductivity sented at the SPE Annual Technical Conference and Exhibition, Denver,
to maximize well performance. The size of the SRV, fracture 21–24 September. doi: 10.2118/115769-MS.
spacing, and azimuth are key components for well-placement Fisher, M.K., Heinze, J.R., Harris, C.D., Davidson, B.M., Wright, C.A.,
and -spacing strategies. and Dunn, K.P. 2004. Optimizing Horizontal Completion Techniques
2. Natural features controlling SRV and fracture spacing in shales in the Barnett Shale Using Microseismic Fracture Mapping. Paper SPE
include shale thickness, stress field and magnitude, presence of 90051 presented at the SPE Annual Technical Conference and Exhibi-
pre-existing open or healed natural fractures, fracture barriers, tion, Houston, 26–29 September. doi: 10.2118/90051-MS.
rock brittleness, and geologic features such as faults and karsts. Fisher, M.K., Wright, C.A., Davidson, B.M., Goodwin, A.K., Fielder,
While these parameters are not under direct human control, it is E.O., Buckler, W.S., and Steinsberger, N.P. 2002. Integrating Fracture

February 2010 SPE Production & Operations 97


Mapping Technologies to Optimize Stimulations in the Barnett Shale. evaluating hydraulic-fracturing treatments and reservoir simu-
Paper SPE 77441 presented at the SPE Annual Technical Conference lation studies, as well as interpreting diagnostic fracture injec-
and Exhibition, San Antonio, Texas, USA, 29 September–2 October. tion test, pressure transient, and production data. Lolon has 8
doi: 10.2118/77441-MS. years of reservoir engineering experience in the petroleum
industry. He previously worked for Pinnacle Technologies as a
Maxwell, S.C., Urbancik, T.I., Steinsberger, N.P., and Zinno, R. 2002.
staff petroleum engineer and at Chevron Pacific Indonesia as
Microseismic Imaging of Hydraulic Fracture Complexity in the Bar- a petroleum engineer. He holds a BS degree in petroleum engi-
nett Shale. Paper SPE 77440 presented at the SPE Annual Technology neering from Bandung Institute of Technology in Indonesia and
Conference and Exhibition, San Antonio, Texas, USA, 29 September–2 MS and Ph.D. degrees in petroleum engineering from Texas
October. doi: 10.2118/77440-MS. A&M University. Norm Warpinski is the director of technology
Maxwell, S.C., Waltman, C.K., Warpinski, N.R., Mayerhofer, M.J., and development for Pinnacle, a Halliburton Service, in Houston. He
Boroumand, N. 2006. Imaging Seismic Deformation Induced by is in charge of developing new tools and analyses for hydrau-
Hydraulic Fracture Complexity. Paper SPE 102801 presented at the lic-fracture mapping, reservoir monitoring, hydraulic-fracture
SPE Annual Technology Conference and Exhibition, San Antonio, design and analysis, and integrated solutions for reservoir devel-
opment. He received his MS and PhD degrees in mechanical
Texas, USA, 24–27 September. doi: 10.2118/102801-MS.
engineering from the University of Illinois, Champaign/Urbana
Mayerhofer, M.J., Lolon, E.P., Youngblood, J.E., and Heinze, J.R. 2006. after receiving a BS degree in mechanical engineering from
Integration of Microseismic Fracture Mapping Results With Numeri- Illinois Institute of Technology. He worked for 28 years for Sandia
cal Fracture Network Production Modeling in the Barnett Shale. National Laboratories performing oil, gas, geothermal, waste
Paper SPE 102103 presented at the SPE Annual Technical Conference repository, and carbon sequestration research before join-
and Exhibition, San Antonio, Texas, USA, 24–27 September. doi: ing Pinnacle in 2005. He served as executive editor of SPE
10.2118/102103-MS. Production & Operations from 2006–08, as review chair of JPT
Rutledge, J.T. and Phillips, W.S. 2003. Hydraulic stimulation of natural from 1990–92, on the SPE R&D committee from 2002–05, as an
fractures as revealed by induced microearthquakes, Carthage Cotton SPE distinguished lecturer in 1998–99, and on various confer-
ence and forum committees. Craig Cipolla is chief engineer-
Valley gas field, East Texas. Geophysics 68 (2): 441. doi: 10.1190/
ing advisor for Schlumberger, Hydraulic Fracture Monitoring,
1.1567212. focusing on the application of microseismic fracture mapping
Warpinski, N.R., Mayerhofer, M.J., Vincent, M.C., Cipolla, C.L., and measurements to improve stimulation designs and field devel-
Lolon, E.P. 2008. Stimulating Unconventional Reservoirs: Maximizing opment. Prior to joining Schlumberger, he was vice president of
Network Growth While Optimizing Fracture Conductivity. Paper SPE stimulation technology for Carbo Ceramics. Prior to the sale of
114173 presented at the SPE Unconventional Reservoirs Conference, Pinnacle Technologies’ fracture mapping and reservoir moni-
Keystone, Colorado, USA, 10–12 February. doi: 10.2118/114173-MS. toring businesses to Halliburton in 2008, Cipolla led Pinnacle’s
Warpinski, N.R., Wolhart, S.L., and Wright, C.A. 2004. Analysis and engineering consulting and fracture mapping applications
Prediction of Microseismicity Induced by Hydraulic Fracturing. SPE J. groups for 12 years. Cipolla remained with Carbo Ceramics-
StrataGen Engineering when the consulting arm of Pinnacle
9 (1): 24–33. SPE-87673-PA. doi: 10.2118/87673-PA.
was split off and renamed in 2008. Craig’s 28 years of experi-
Zimmer, U., Maxwell, S., Waltman, C., and Warpinski, N. 2007. Microseis- ence includes the design and evaluation of hydraulic-fractur-
mic Monitoring Quality Control (QC) Reports as an Interpretive Tool ing treatments, training engineers in the use of real-time data
for Nonspecialists. Paper SPE 110517 presented at the SPE Annual analysis, tight gas reservoir engineering, integrated field stud-
Technical Conference and Exhibition, Anaheim, California, USA, ies, and supervising stimulation treatments. In addition, He is an
11–14 November. doi: 10.2118/110517-MS. expert in the application of microseismic and tiltmeter-fracture
mapping technologies. Prior to joining Pinnacle in 1996, Cipolla
held positions with Union Pacific Resources, CER Corporation,
SI Metric Conversion Factors and Dresser Titan. Doug Wasler has worked for Pinnacle for the
past five years. He has 30 years of experience in the Permian
acre × 4.046 873 E+03 = m2 Basin, mid-continent, Appalachia, Rockies, and South Texas with
bbl × 1.589 874 E–01 = m3 Dowell Schlumberger, The Western Company of North America,
cp × 1.0* E–03 = Pa·s BJ Services, and Pinnacle. He specializes in the calibration of
3D fracture modeling through a number of methods, includ-
ft × 3.048* E–01 = m
ing historical production transient analysis and calibration by
*Conversion factor is exact. various fracture mapping processes. He has taught numerous
seminars and short courses on subjects related to his fields of
interest, including Mini-frac Evaluation, 3-D Pressure Matching,
Michael Mayerhofer is the technology manager for fracture On-Location Stimulation QC, Fracturing Situations, Horizontal
diagnostics at Pinnacle, a Halliburton Service in Houston. His Devonian Completion Best Practices, Fracture Mapping
responsibilities include the application of tiltmeter and micro- Techniques, and others. Most recently, he has specialized in
seismic hydraulic-fracture-mapping results for optimizing frac- the examination and comparison of the various emerging
ture completion, well placement and infill drilling strategies, resource plays in North America. He has authored 14 papers
the design and evaluation of hydraulic-fracturing treatments, and holds three patents in his areas of interest. He holds a BS
reservoir engineering, and integrated field studies. His 18-year degree in natural gas engineering from Texas A&M University.
involvement with hydraulic fracturing and reservoir engineer- Mike Rightmire is a senior vice president with Forrest A. Garb
ing includes fundamental research and real-field applications and Associates where he is responsible for reservoir engineer-
in various global producing areas and has resulted in more ing projects and reserves appraisals. During his 25-year career,
than 40 technical papers and journal articles. Prior to joining Rightmire also worked for Arco and Pinnacle Technologies. His
Pinnacle Technologies in 1997, he worked for Union Pacific tenure at Arco included experience in reservoir and produc-
Resources in Fort Worth, Texas, USA. He holds a PhD degree in tion engineering, software development, numerical simula-
petroleum engineering from Mining University Leoben in Austria. tions, chemical transport modeling, and reservoir engineering
He was a member of the SPE Well Completions committee from training. Responsibilities with Pinnacle Technologies included
1998 to 2001 and has served on the JPT editorial committee. reservoir surveillance applications and project management.
Elyezer Lolon is a senior staff engineer at StrataGen Engineering Rightmire holds a BS degree in petroleum engineering from
(CARBO Ceramics) in Houston. e-mail: elyezer.lolon@strata- Texas A&M University and a BS degree in biology from the
genengineering.com. His responsibilities include designing and University of Alaska.

98 February 2010 SPE Production & Operations


Modeling of Hydraulic-Fracture-Network
Propagation in a Naturally Fractured
Formation
X. Weng, O. Kresse, C. Cohen, R. Wu, and H. Gu, Schlumberger

Summary of the fracture network, as compared to biwing fractures, and is


Hydraulic fracturing in shale-gas reservoirs has often resulted in thus able to predict a reasonable network footprint. This is the
complex-fracture-network growth, as evidenced by microseismic case provided that the assumed fracture spacings and stress anisot-
monitoring. The nature and degree of fracture complexity must be ropy (defined as the difference between maximum and minimum
understood clearly to optimize stimulation design and completion horizontal stresses) have been calibrated properly against micro-
strategy. Unfortunately, the existing single-planar-fracture models seismic observations. Even though the microseismic-event cloud
used in the industry today are not able to simulate complex frac- does not provide the precision to delineate the exact boundary of
ture networks. the stimulated volume and there are inherent uncertainties in the
A new hydraulic-fracture model is developed to simulate complex- estimated dimensions, it provides engineers with an independent
fracture-network propagation in a formation with pre-existing measurement to validate the assumed fracture spacings and to
natural fractures. The model solves a system of equations govern- make adjustments if needed because these parameters are highly
ing fracture deformation, height growth, fluid flow, and proppant uncertain.
transport in a complex fracture network with multiple propagat- While the wiremesh model is a significant improvement over
ing fracture tips. The interaction between a hydraulic fracture and the conventional biwing-fracture model in terms of being able to
pre-existing natural fractures is taken into account by using an simulate complex fracturing in shale gas, and it provides an esti-
analytical crossing model and is validated against experimental mate of the fracture-network dimensions and proppant placement
data. The model is able to predict whether a hydraulic-fracture in the network, it has some limitations. One of the limitations
front crosses or is arrested by a natural fracture it encounters, which is that the fracture-network pattern (i.e., the fracture spacings)
leads to complexity. It also considers the mechanical interaction cannot be linked directly to the pre-existing natural fractures.
among the adjacent fractures (i.e., the “stress shadow” effect). An Calibration against microseismic data is required to obtain frac-
efficient numerical scheme is used in the model so it can simulate ture spacings consistent with the stimulated volume indicated by
the complex problem in a relatively short computation time to the microseismic data. However, the calibrated fracture spacings
allow for day-to-day engineering design use. cannot be used easily in other wells, or even other stages in the
Simulation results from the new complex-fracture model show same well, if the reservoir characteristics and pumping param-
that stress anisotropy, natural fractures, and interfacial friction play eters change significantly. When used for parametric studies to
critical roles in creating fracture-network complexity. Decreasing optimize a treatment, the outcome may be skewed because the
stress anisotropy or interfacial friction can change the induced- model assumes fixed spacings, and any effect of treating param-
fracture geometry from a biwing fracture to a complex fracture eters on the resulting fracture-network pattern is not taken into
network for the same initial natural fractures. The results presented account. Another limitation is that network geometry is assumed
illustrate the importance of rock fabrics and stresses on fracture symmetric with respect to the injection point, and the shape is
complexity in unconventional reservoirs. These results have major elliptical, which limits the ability to match microseismic data that
implications for matching microseismic observations and improv- show significant asymmetry or irregular shape in the stimulated
ing fracture stimulation design. region.
A more rigorous general hydraulic-fracture model incorporat-
Introduction ing a full solution of coupled elasticity and fluid equations has been
presented by Zhang et al. (2007), but it is limited to the 2D plane-
Hydraulic fracturing in shale gas reservoirs has often resulted in strain conditions, which is suitable for situations where fracture
complex-fracture-network growth, as evidenced by microseismic length is much greater than height but is not suitable for general
monitoring (Fisher et al. 2002; Maxwell et al. 2002; Daniels et al. field applications where fractures are often contained in height.
2007; Le Calvez et al. 2007). This renders the hydraulic-fracture Additionally, the model is computationally intensive and not well
models used in the industry today for simulating a biwing planar suited for engineering design applications. Olson (2008) presented
fracture in a conventional reservoir unsuitable for designing frac- a complex-fracture-network model that is capable of predicting
ture treatments in a shale-gas formation where complex fracturing hydraulic-fracture propagation and interaction with pre-existing
is evident. To overcome the limitation of the conventional fracture natural fractures. However, the model is based on only fracture
models, hydraulic-fracture models for simulating propagation of mechanics and does not include fluid flow and proppant transport.
fracture networks consisting of two orthogonal sets of parallel and A constant fracturing pressure is assumed in the simulation. This
uniformly spaced fractures (referred to as a “wiremesh” model in limits its ability to accurately simulate actual hydraulic-fracture
this paper) have been developed (Xu et al. 2009; Xu et al. 2010; treatments.
Meyer and Bazan 2011). Although the assumed fracture geometry In this paper, a new hydraulic-fracture model is presented that
may not accurately represent the actual complex fracture network, simulates propagation of complex fractures in a formation with
which may consist of irregularly oriented and distributed fractures, pre-existing natural fractures. The model solves a system of equa-
it is a reasonable approximation because it properly accounts for tions governing fracture deformation, height growth, fluid flow,
the increased storage and surface area and mechanical interaction and proppant transport in a complex fracture network with mul-
tiple propagating fracture tips. The interaction between hydraulic
fractures and pre-existing natural fractures is taken into account by
Copyright © 2011 Society of Petroleum Engineers
use of an analytical crossing model. Numerical-simulation results
This paper (SPE 140253) was accepted for presentation at the SPE Hydraulic Fracturing and the impact of natural-fracture distribution, interfacial friction,
Technology Conference, The Woodlands, Texas, USA, 24–26 January 2011, and revised for
publication. Original manuscript received 10 February 2010. Revised manuscript received
and stress anisotropy on fracture pattern are illustrated through a
11 July 2011. Paper peer approved 9 September 2011. simple example.

368 November 2011 SPE Production & Operations


illustrates the hydraulic-fracture network created by the fracturing
fluid in a formation with pre-existing natural fractures.
The equation for flow of a power-law fluid in any given fracture
branch can be expressed using the Poiseuille law (Nolte 1991):
Hydraulic frac 2 n ′+1
n′
2 K ′ ⎛ 4 n′ + 2 ⎞ 1 ⎛ w( z ) ⎞ n′
␣0 = n' ⎜ ⎟ ; ␾ ( n′ ) = ∫ ⎜⎝ ⎟ dz
␾ ( n′ ) ⎝ n′ ⎠ H fl H fl
w ⎠

n ′−1
∂p 1 q q
= −␣0 2 n′+1 , . . . . . . . . . . . . . . . . . . . . . . . . (1)
∂s w H fl H fl

where p is fluid pressure; q is the local flow rate in the fracture;


Hfl is the height of the fluid in the fracture; w(z) is fracture width
Natural frac σh
as a function of depth z; w is the average width; n' and K' are fluid
power-law index and consistency index, respectively; and s is the
σH distance along the fracture.
The local condition for mass balance is given by the continu-
ity equation:
Fig. 1—Diagram of hydraulic-fracture network and pre-existing
natural fractures. ∂q ∂( H fl w) 2hLC L
+ + qL = 0, qL = , t > ␶ 0 (s) , . . . . . . . (2)
∂s ∂t t − ␶ 0 (s)
Unconventional Fracture Model where CL is the total leakoff coefficient, hL is the height of leakoff
The complex-fracture-network model, referred to as an uncon- zone, and ␶ 0 (s) is the time when each element of a fracture is
ventional fracture model (UFM) in this paper, simulates the first exposed to fracturing fluid. Also, the global volume balance
propagation, deformation, and fluid flow in a complex network must be satisfied:
of fractures. The model solves the fully coupled problem of fluid t L (t ) L (t ) t

flow in the fracture network and the elastic deformation of the ∫ Q (t ) d t = ∫ Hw d s + ∫ ∫q L d t d s , . . . . . . . . . . . . . . . . . (3)
fractures, which has similar assumptions and governing equations 0 0 0 0

to those of conventional pseudo-3D fracture models. But instead where Q(t) is pump rate, L(t) is the total length of all fracture
of solving the problem for a single planar fracture, UFM solves branches at time t, and H(s, t) is fracture height. Additionally, at
the equations for the complex fracture network. Fracture-height any junction of fracture branches, the sum of flow rates into the
growth is modeled in the same manner as in conventional pseudo- junction from all branches must be equal to zero. At the wellbore,
3D models. A three-layer proppant-transport model, consisting of the sum of flow rates into all open perforation intervals must be
a proppant bank at the bottom, a slurry layer in the middle, and equal to the prescribed injection rate Q(t); i.e.,
clean fluid at the top, is adopted for simulating proppant transport
in the fracture network. Transport equations are solved for each ∑ q (t ) = Q(t ),
i i = 1,.. N perf , . . . . . . . . . . . . . . . . . . . . . . . . (4)
component of the fluids and proppants pumped. A key difference i

between UFM and the conventional planar-fracture model is being where qi(t) is the injection rate into each perforation cluster (i =
able to simulate the interaction of hydraulic fractures with pre- 1,..Nperf) and is determined as a part of the solution.
existing natural fractures (i.e., determining whether a hydraulic Eqs. 1 and 2 provide the expressions for solving pressure and
fracture propagates through or is arrested by a natural fracture flow distribution in the fracture network for a given fracture width
when they intersect and whether it subsequently propagates along distribution. On the other hand, fracture-width is related to fluid
the natural fracture). The branching of the hydraulic fracture at the pressure through the elasticity equation. To have a computationally
intersection with the natural fracture gives rise to the development efficient model for general engineering design applications, the
of a nonplanar, complex fracture pattern. A crossing model that is same assumption of 2D plane-strain solution for fracture width
extended from the Renshaw and Pollard (1995) interface crossing as in conventional pseudo-3D models is adopted. For a vertical
criterion, applicable to any intersection angle, has been developed fracture in a layered medium with variable minimum horizontal
(Gu and Weng 2010) and validated against the experimental data stress σh(x, y, z) and fluid pressure p, the width profile can be
(Gu et al. 2011), and it is integrated in the UFM. In addition determined from an analytical solution given as (refer to Eq. 9
to the hydraulic-fracture/natural-fracture interactions, the UFM below for detailed form of solution)
also takes into account the interaction among adjacent hydraulic-
fracture branches by computing the stress-shadow effect on each w( x , y, z ) = w( p( x , y), H , z ). . . . . . . . . . . . . . . . . . . . . . . . . . (5)
fracture by the adjacent fractures.
In the limiting case of a fracture in a zone of uniform in-situ
Governing Equations. To simulate the propagation of a complex stress and constant height, the width/pressure relation in Eq. 5
fracture network that consists of many intersecting fractures, the reduces to that of a 2D PKN model (here E' is the plane-strain
equations express the underlying physics of the fracturing process Young’s modulus):
must be satisfied. These include the equation expressing the fluid
flow in the fracture network, the equation expressing the fracture ␲ H( p − ␴n )
deformation, and the fracture-propagation criterion. In the most w= . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
2E '
general sense, hydraulic fractures are 3D in nature, and the fracture
planes may not be vertical. This is especially the case when the At fracture tips, the following boundary conditions are
initial natural fractures are not vertical, potentially leading to non- satisfied:
vertical hydraulic fractures when the fracturing fluid opens up the
natural fractures. However, solving a fully 3D nonplanar fracture p = ␴ n , W = 0, q = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
problem is extremely computation intensive and not suitable for
engineering design models. In the model presented in this paper, Eqs. 1 through 7 represent the complete set of governing
we make the assumption that all natural and hydraulic fractures equations for the problem and must be solved at each timestep to
are vertical. Therefore, the model is applicable only to a formation find fracture opening, fluid pressure, and flow rate in the entire
with vertical or near-vertical natural fractures. Fig. 1 schematically fracture network.

November 2011 SPE Production & Operations 369


At the first timestep, it is assumed that biwing fractures are initi- Fracture Interaction. Fracture interaction is one of the most
ated from all perforation intervals in the lowest-stress layer. A 2D important factors that must be taken into account to model hydrau-
analytical solution for power-law fluid is used to calculate initial fluid lic-fracture propagation in naturally fractured reservoirs. This
pressure for a prescribed initial fracture length L0 at a known injection includes the interaction between hydraulic fractures and natural
rate Q0. The initial time t0 can be found from volume balance. fractures, as well as between hydraulic fractures.
The interaction between hydraulic and natural fractures is
Height Growth. In a multilayered formation, the fracture height influenced by the natural fractures’ orientation and geometry, in-
and width profile of a fracture cross section depends on fluid situ stresses, rock properties (e.g., rock modulus, tensile strength,
pressure, the in-situ stress, fracture toughness, layer thickness, toughness, permeability, and porosity), fluid properties (e.g., pres-
and elastic modulus of each layer covered by the fracture height. sure and rheology), and interface properties (e.g., friction coef-
Calculation of height growth takes the same approach as that used ficient, cohesion, and permeability), and it represents a coupled
for cell-based pseudo-3D models (Mack and Warpinski 2000). rock-deformation/frictional-sliding/fluid-flow problem. It is in
Each element (equivalent to a “cell” in the cell-based pseudo-3D itself a very complex process to model.
models) has its own height. The height and width profile of the The behavior of a hydraulic fracture when it intersects natural
fracture cross section is determined only on the basis of local fractures (crossing, arresting, branching, or reinitiating with an off-
pressure and vertical stress profile. Interaction between adjacent set) is one of the root causes of the creation of complex fractures.
cells is neglected. If the hydraulic fracture crosses a natural fracture, the hydraulic
Similar to a pseudo-3D model, the fluid flow and the associated fracture remains planar, and there is no complexity generated as
pressure gradient in the vertical direction are negligible except near long as the fluid pressure does not subsequently exceed the open-
the upper- and lower-fracture-tip region. That is, ing pressure of the natural fracture. Conversely, if the hydraulic
fracture does not cross the natural fracture, but instead opens the
natural fracture and propagates it along, a complex fracture net-
p = pcp + ␳ f g(hcp − z ), . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
work may form. It is important to determine whether a hydraulic
fracture crosses natural fractures under particular field conditions
where pcp is the pressure in the fracture at a reference depth hcp of in-situ stress and rock and natural-fracture properties.
(chosen as the perforation depth) measured from the bottom tip, z Theoretical, numerical, and experimental works have been
is the position in the fracture, and ␳f is the fluid density. conducted by various researchers on fracture propagation across
The stress-intensity factors at the fracture top and bottom tips interfaces (Renshaw and Pollard 1995; Blanton 1982; Warpinski
are calculated from the pressure inside the fracture, the fracture and Teufel 1987; Thiercelin and Makkhyu 2007; Zhang and Jeffrey
geometry (height, in this case), and the layer stresses. A stable frac- 2006). Warpinski and Teufel (1987) observed in their mineback
ture height is determined by matching the stress-intensity factors experiments that the hydraulic fracture was often arrested at the
at the tips to the fracture toughness in the layers that contain the rock joints, leading to branching of multiple fractures and a com-
fracture tips. The relations for the stress-intensity factors and the plex-fracture pattern. Some of the mechanisms have been studied,
width, as functions of pressure and tip positions (fracture height), and component models for the mechanisms were proposed. Ren-
are given in Mack and Warpinski (2000): shaw and Pollard (1995) developed a simple criterion for predict-
ing whether a fracture will propagate across a frictional interface
␲h ⎡ ⎛ 3 ⎞⎤ orthogonal to the approaching fracture. The criterion is based on
K Iu = pcp − ␴ n + ␳ f g ⎜ hcp − h ⎟ ⎥
2 ⎢⎣ ⎝ 4 ⎠⎦ the linear elastic fracture mechanics solution for the stresses near
the fracture tip to determine the stresses required to prevent slip
2 n −1 ⎡h ⎛ h − 2h ⎞ ⎤
+ ∑ (␴ − ␴ i ) ⎢⎣ 2 arccos ⎜⎝ h i ⎟⎠ − hi (h − hi ) ⎥⎦ ,
␲ h i =1 i +1
along the interface at the moment when the stress on the opposite
side of the interface is sufficient to reinitiate a fracture. Compres-
sional crossing will occur if the magnitude of the compression act-
ing perpendicular to the frictional interface is sufficient to prevent
␲h ⎡ ⎛ h⎞ ⎤ slip along the interface at the moment when the stress ahead of the
K Il = pcp − ␴ n + ␳ f g ⎜ hcp − ⎟ ⎥
2 ⎢⎣ ⎝ 4⎠ ⎦ fracture tip is sufficient to initiate a fracture on the opposite side
2 n −1 of the interface. The criterion is given as
⎡h ⎛ h − 2h ⎞ ⎤
+ ∑ (␴ − ␴ i ) ⎢⎣ 2 arccos ⎜⎝ h i ⎟⎠ + hi (h − hi ) ⎥⎦ ,
␲ h i =1 i +1 0.35
0.35 +
−␴ xx kfric
>
and w( z ) =
4
( )
⎡ pcp − ␴ n + ␳ f g hcp − z ⎤ z (h − z )
E′ ⎣ ⎦ T0 − ␴ yy 1.06
, . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)

⎡ ⎛ h − 2hi ⎞ ⎤
⎢ z⎜ ⎟⎠ + hi ⎥ where ␴xx and ␴yy are far-field effective stresses (tensile is positive)
− ⎝ h
4 n −1 ⎢(hi − z ) cosh 1
⎥ acting parallel with and perpendicular to a propagating hydraulic
+ ∑
␲ E ′ i =1
(␴ i +1 − ␴ i ) ⎢ z − hi ⎥, fracture correspondingly, T0 is rock tensile strength, and kfric is a
⎢ − ⎥ coefficient of friction at the interface. This criterion was validated
⎢ + z (hh − z ) arccos ⎜⎛ h 2 hi ⎞ ⎥
⎝ h ⎠ ⎥⎦ ⎟ by laboratory experiments.
⎢⎣
Natural fractures are often not aligned with the contemporary
. . . . . . . . . . . . . . . . . . . . . . . . (9) principal in-situ-stress directions, in which case the intersection
angle of a hydraulic fracture approaching a natural fracture is
where KIu and KIl are the stress intensity-factors at the upper and between 0 and 90° (angle ␤ in Fig. 2). Therefore, it is important
lower tips, respectively; w(z) is the width at depth z; ␴n and ␴i are to extend the Renshaw and Pollard criterion to nonorthogonal
the in-situ stresses at the top tip and the ith layer, respectively; h intersections.
is the fracture height; and hi represents the height from the lower The Renshaw and Pollard model recently has been extended
tip to the top of the ith layer. to nonorthogonal intersections by Gu and Weng (2010). Because
This model, called the equilibrium height model, is extended the model is based on an analytical solution, it is efficient in com-
further to a nonequilibrium height model, in which the pressure putation and suitable for incorporation in the hydraulic-fracture-
gradient because of the fluid flow in the tip regions in the verti- network model described in this paper.
cal direction is taken into account. This adds an apparent fracture For hydraulic-fracture applications, the ratio of the remote
toughness proportional to the fracture’s top and bottom velocities in-situ stresses (␴H to ␴h) is larger than unity. The coefficient of
in the equations above. The resulting equations are solved numeri- friction for many rocks is typically between 0.1 and 0.9. Fig. 3
cally to determine the tip velocities. shows the results from the extended criterion for a number of

370 November 2011 SPE Production & Operations


10
h
βx 90°

Stress ratio (hmax / hmin)


y 75°
60°
interface 45°
30°
15°
x
y
H

1
βy 0 0.2 0.4 0.6 0.8 1
fracture x Coefficient of friction

Fig. 3—Plot of crossing criterion for stress ratio > 1 at several


intersection angles (for T0 = 0).

because of the stress-shadow effects is computed at each timestep in


the UFM model and is then added to the initial in-situ-stress field
on each fracture element during pressure and width iteration. At the
Fig. 2—Schematic of fracture approaching an interface.
time of writing of this paper, only the contribution of the normal dis-
placement (i.e., fracture-opening width) is considered. The effect of
intersection angles. The region to the right of each curve represents stress shadow, especially the shear stress, on the directional change
the crossing condition. It can be seen that as the angle decreases of propagating fracture tips is not considered at present.
from 90°, crossing becomes increasingly difficult. In other words,
a hydraulic fracture is more likely to be arrested and propagate Proppant Transport. Proppant transport is an important feature
along the natural fracture it intersects. The extended criterion also of the modeling work involved in the development of a hydraulic-
has been compared to laboratory experimental results and showed fracturing simulator because it determines the final conductivity
good agreement (Gu et al. 2011). of the created fracture network. The model presented in this
paper considers the transport of multiple fluid and proppant spe-
Stress Shadow. Fracture-network-growth pattern is affected by cies inside the fracture network. It accounts for several physical
the mechanical interaction among the adjacent fractures. Generally mechanisms such as bridging, packing, and the settling of proppant
known as the stress-shadow effect, the stress field of one fracture to form a bank, as well as the erosion of the existing bank. To
is perturbed by the opening and shearing displacements of other achieve computational efficiency, a 1D proppant-transport model
nearby fractures. In a 2D, plane strain, displacement discontinuity in the horizontal direction has been adopted, while the vertical
solution, Crouch and Starfield (1983) described the normal and flow of proppant as a consequence of both settling and bank ero-
shear stresses (␴n and ␴s) acting on one fracture element induced sion is computed to track the height of the bank, the slurry, and
by the opening and shearing displacement discontinuities (Dn and the clean fluid locally.
Ds) of all fracture elements as the following: The fluid and proppant species are described in the model
through volumetric concentrations. They are calculated in the trans-
␴ ni = ∑ j =1 Cnsij Dsj + ∑ j =1 Cnn
N N ij
Dnj, port algorithm as averaged values over the volume inside the frac-
ture element, from the top of the bank to the top of the fracture:
and ␴ si = ∑ j =1 Cnsij Dsj + ∑ j =1 Cnn
N N ij
Dnj . . . . . . . . . . . . . . . . . (11) w ′ ⌬x ′
xc +
H 2 2
1
X k ( x ′, y′, z ) d x d y′ d z ,
⌬x ′w ( H − H bank ) H∫bank ∫ ∫
ck =
where Cij are the 2D, plane-strain elastic influence coefficients, w ′ ⌬x ′
and their expressions can be found in Crouch and Starfield (1983). − xc −
2 2
This method is used in our model to compute the additional stress . . . . . . . . . . . . . . . . . . . . . (12)
induced on each fracture element from the displacements of adja-
cent elements (Fig. 4). In addition, a 3D correction factor suggested where Xk is the volume fraction of a species identified by the index
by Olson (2004) was further introduced to the influence coefficients k, Δx' is the length of the element, H is the height of the fracture,
(Cij in Eq. 11) to account for the 3D effect caused by finite fracture Hbank is the height of the bank, and ck is the concentration of species
height that leads to decaying of interaction between any two fracture k (fluid or proppant). The 1D transport model inside the fracture is
elements when the distance increases. This additional normal stress similar to the one in Adachi et al. (2007) except that it considers
only advective transport in the horizontal direction. For example,
Dnj Eq. 13 describes the variation of fluid concentration cfl,k as governed
by the flow inside the fracture and the leakoff into the matrix:

Dsj
∂ ( H − H bank ) wc fl ,k
+
(
∂ q fl c fl ,k )=−f c fl ,k , . . . . . . . . . . (13)
∂t ∂x
Elem j leakoff

where qfl is the flow rate (m3/s) and fleakoff is the flux of fluid through
σsi Elem i
the fracture walls (m2/s) divided by the fracture height above the
bank. The equation is similar in the case of proppant species,
σni
except that there is no proppant leakoff into the matrix. For the
calculation of slurry height, a similar transport equation is used to
calculate the volumetric concentration of the slurry. The numerical
Fig. 4—2D displacement discontinuity method (DDM) approach: implementation of these transport equations uses an explicit
influence of displacement discontinuities from jth element on scheme in time and a total variation diminishing method for the
stress of ith element. advection term (LeVeque 1992).

November 2011 SPE Production & Operations 371


( )
1/ n ′
New guessed ⎡ 1 ␳ prop,k − ␳ fl ⎤
time step vset,k = ⎢ n′−1 gDkn′+1 ⎥ , . . . . . . . . . . . . . . . . (14)
⎢⎣ 3 18 K′ ⎥⎦

where vset,k is the settling velocity for proppant type k, ␳prop,k is its
Extend all tips by density, and Dk is its diameter. K' and n' are the averaged flow con-
ΔLi based on velocity sistency index and averaged flow behavior index, respectively, both
weighted by the fluid concentration. A more accurate calculation
of the settling velocity has also been implemented, on the basis of
the correlation from Shiller and Naumann (1933), to calculate the
Apply junction/ drag-force coefficient on a spherical particle.
crossing rules The modeling of erosion of the bank is based on the correlation
from Wang et al. (2003), which describes the minimal height of
clean fluid and slurry at given flow conditions, under which the
proppant does not settle and the bank is eroded.
Solve coupled flow
Numerical Solution. At any given timestep, the created hydraulic-
and elasticity
Eqs. for p, q, w
fracture network is represented by connected fracture elements.
New elements are added as fracture tips advance. Many new
Iterate elements can be added at all propagating fracture tips in a single
timestep, and new element numbers are assigned to these new
elements. As a result, the element numbers are not continuous
Update Δt based from one element to the next. Therefore, a special code structure
on mass balance is established to track the neighbors, and special junction elements
at intersection of multibranches are also introduced.
The combined Poiseuille and continuity equations (Eqs. 1
and 2) are discretized in terms of pressure pi and average width
Apply proppant transport wi defined at the center of each element and flow rate qij defined
at the boundary between Element i and Element j. The resulting
mass-balance equation is written as
Update fracture height
d Volfluid (i, t ) = d Volfrac (i, t ) + d Volleakoff (i, t ), . . . . . . . . . . . . (15)

or, in the form of pressure, width, and flow rate at each element,
Compute stress shadow

− ∑ qijt d t = (h tfl ,i wit lit − h tfl−,idt wit − dt lit − dt )


Fig. 5—Structure of propagation-solution loop. j

h ( k , i ) ⎡ hocc ( k , i )li t − ␶ 0 ( k, i ) ⎤
t t
N zones

+ 2 ∑ ⎨2C L ( k , i ) L ⎢ ⎥
k =1 ⎩ h( k , i ) ⎢ − hocc ( k , i )li
t − dt t − d t
t − ␶ 0 ( k , i ) ⎥⎦
old

In the vertical direction, the model for proppant considers two hL ( k , i ) t ⎫
+ S p (k, i) ⎡ h ( k , i )lit − h t − d t ( k , i )lit − d t ⎤⎦ ⎬ . . . . . . . . . . . (16)
h( k , i ) ⎣
competing mechanisms: the proppant settling into the bank and the
erosion of the existing bank. The settling is calculated by first cal- ⎭
culating the settling velocity for each proppant and then calculating
the increase of the bank height on the basis of both the proppant Eq. 16 includes the leakoff coefficient CL (k, i) for each Zone k
concentration and the height of the slurry. The settling velocity for covered by Fracture Element i, as well as the height of each zone
each proppant was calculated initially on the basis of the extension h (k, i), leakoff height of each zone hL (k, i), the height of the
t
of the Stokes’s law to power fluids given by Daneshy (1978): occupied part of the zone hocc ( k , i ) at position of Element i, and

Injection Pressure vs. Time Fracture Half-Length vs. Time


1500
33600
1300
33500
Fracture Length (m)
Inj. Pressure (KPa)

33400 1100
33300 900
33200
700
33100
500
33000
UFM
32900 Conventional bi-wing 300 UFM
32800 Conventional bi-wing
100
32700
0 2000 4000 6000 8000 –100
0 1000 2000 3000 4000 5000 6000 7000 8000
Time (sec)
Pumping Time (sec)

Fig. 6—Comparison of injection pressure and fracture half-length between UFM and a conventional model for a biwing fracture.

372 November 2011 SPE Production & Operations


terms of pressure by using a damped Newton-Raphson method.
TABLE 1—INPUT PARAMETERS FOR THE BIWING CASE
The Newton-Raphson method (or Newton method) is a popular and
3
Injection rate 0.5 m /s powerful method that is applicable to solution of multidimensional
Minimum confining stress 4.8e+07 Pa nonlinear systems of equations. It is based on the simple first-
Young’s modulus 3.5e+10 Pa
order Taylor-series expansion. For any given approximate solution
p, f(p) ≠ 0; otherwise, p is already a solution, an improved solution
Fracture height 10 m p + δp is sought after such that f(p + δp) ≈ f(p) + Jf(p)δp = 0. Here
Viscosity 0.001 kg/m.s Jf(p) is the Jacobian matrix for f, defined as

the spurt Sp (k, i). Fluid-flow rate between adjacent elements is ⎡⎣ J f ( p ) ⎤⎦ =


( ).
∂fi p kj
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (18)
calculated by integrating the Poiseuille equation (Eq. 1) and the ij ∂p j
elasticity equation (Eq. 5). For example, in the limiting case of
constant fracture height, it has the form { }
If ␦ p satisfies the linear system,

qij = ␥ ij ⋅ ␤ij 1/ n '


, ␥ ij =
h fl { } { ( )}
⎡⎣ J f ( p ) ⎤⎦ ␦ p j = − fi p kj . . . . . . . . . . . . . . . . . . . . . . . . (19)
ij

(␣ ⋅ (l + l ) ⋅ ( n′ + 1))
1/ n ′
i j Then (p + δp) is taken as an approximate zero of f, or solution
of f(p) = 0. In this sense, Newton’s method replaces a system of
␤ij = ( pi − ␴ i ) ( ) ( ) ( )
2 n ′+ 2 2 n ′+ 2 2 n ′+ 2 2 n ′+ 2
− p* − ␴ i + p* − ␴ j − pj − ␴ j , nonlinear equations with a system of linear equations that must be
solved repeatedly to converge on an adequate solution. In this way,
pi l j + p j li as it iterates, the solution is updated as
p* = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (17)
li + l j
p kj +1 = p kj + ␤␦ p kj +1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20)
where p* is an averaged value of the pressure at the interface
between Elements i and j. Convergence is achieved when some norm associated with
Different forms of flow-rate equation are used and implemented the right-hand side of Eq. 19 (i.e., solution) reaches the specified
in the model for the cases of laminar, turbulent, and Darcy fluid minimum value.
flow. A combination of the above equations leads to a nonlinear After the solution converges, a new timestep starts, and an
system of equations, simplified as f(p) = 0, which is solved in explicit method is used to extend the fracture tips. The algorithm

X-axis
–3200 –2900 –2400 –2000 –1000 –1200 –800 –400 0 400

1400 1400

1200 1200

1000 1000
NF 2
800 800
NF 1
600 600
H perfs
400 400
NF 3 h
200 200

0 Treatment Well 0

–2900 –2400 –2000 –1000 –1200 –800 –400 0 400


X-axis

Fig. 7—Top view of wellbore, perforation, and natural fractures (NF) (x- and y-coordinate in feet).

TABLE 2—INPUT PARAMETERS FOR EXAMPLE NO. 1


Zone Top Min Horiz Young’s Poisson’s
Depth (ft) Thickness (ft) Stress (psi) Modulus (psi) Ratio Permeability (md)

6700 200 5338 4.0e6 0.24 0.0001


6900 300 4538 4.0e6 0.24 0.0001
7300 200 5537 4.0e6 0.24 0.0001
Injection rate 50 bpm
Fluid viscosity 5 cp
Rock tensile strength 500 psi

November 2011 SPE Production & Operations 373


X-axis
Width
–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600
0.5
0.45
0.4
0.35
0.3
0.25
0.2 –2000 X-axis
0.15
–1600
0.1 –1200
0.05 –800
0

–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600


X-axis

(a) t=1 minute

X-axis
Width
–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600
0.5
0.45
0.4
0.35
0.3
0.25
0.2 –2000 X-axis
0.15
–1600
0.1 –1200
0.05 –800
0

–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600


X-axis

(b) t=15 minutes 19 seconds

Figs. 8a and b—Fracture-geometry evolution for Example No. 1 case 1 (x- and y-coordinate in feet).

first finds all the hydraulic-fracture tips that satisfy the propagation A special tip element is implemented, and the pressure at the
criterion (stress-intensity factor is greater than rock toughness). For propagating tip element is prescribed by using asymptotic PKN tip
each propagating tip element, the flow rate into the tip is calculated solution as a function of the element’s length and velocity.
on the basis of the Poiseuille equation (Eq. 1) to find the corre- Fig. 5 shows the general numerical scheme at each timestep.
sponding tip velocity (equal to fluid velocity near the tip):
Model Validation and Examples
vtip = qtip / (h fl w). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (21) Comparison With Conventional Model for a Biwing Fracture.
To check the UFM validity, it is compared with the limiting case
The tip with the highest velocity is extended by a prescribed of biwing fractures. Fig. 6 shows the comparison of UFM results
maximum length lfix. The other tips are extended proportionally to shown as a red curve with a conventional hydraulic-fracture model
their velocities: shown as a blue curve (Gulrajani et al. 1997), for the input param-
eters as given in Table 1. The predicted injection pressure and
ltip = lfix vtip / vtip
max
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (22) fracture length from the two models are identical.

374 November 2011 SPE Production & Operations


X-axis
Width
–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600
0.5
0.45
0.4
0.35
0.3
0.25
0.2 –2000 X-axis
0.15
–1600
0.1 –1200
0.05 –800
0

–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600


X-axis

(c) t=18 minutes 6 seconds

X-axis
Width
–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600
0.5
0.45
0.4
0.35
0.3
0.25
0.2 –2000 X-axis
0.15
–1600
0.1 –1200
0.05 –800
0

–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600


X-axis

(d) t=24 minutes 59 seconds

Figs. 8c and d—Fracture-geometry evolution for Example No. 1 case 1 (x- and y-coordinate in feet).

Example No. 1: Simple Natural Fractures. A major advantage only a few natural fractures is given for different values of these
of the UFM is to be able to predict the impact of natural fractures parameters and different intersection angles. Fig. 7 shows the top
and in-situ stresses on the created hydraulic-fracture network. The view of the wellbore and a single perforation cluster where the
factors that influence fracture complexity include natural-frac- fracture is initiated and the prescribed natural fractures. The input
ture size, orientation, and distribution; natural-fracture mechanical parameters for the example are given in Table 2.
properties (interfacial friction coefficient, cohesion, and fracture Case 1: Friction Coefficient = 0.5, Stress Anisotropy = 200
toughness); stress anisotropy (defined as difference between maxi- psi, Intersection Angle = 90°. In this example case, the property of
mum and minimum horizontal stresses); and rock tensile strength. NF 3 is chosen so that a hydraulic fracture cannot propagate along
As shown in Fig. 4, whether a hydraulic fracture crosses a natural NF 3 easily (a large toughness is used). This biases the fracture
fracture when they intersect depends largely on the interfacial propagation to the right side of the wellbore, so the fracture geom-
friction coefficient and stress anisotropy. Therefore, these two etry does not become very complex. Fig. 8 shows the simulation
parameters play a critical role in fracture complexity. fracture geometry at different times, containing top views of both
To illustrate the importance of the natural friction coefficient the hydraulic fracture and natural fractures and 3D views with
and stress anisotropy, a simple example involving a rock containing color contour of width for the hydraulic fracture only.

November 2011 SPE Production & Operations 375


X-axis Width
–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600
0.5
0.45
0.4
0.35
0.3
0.25
0.2 –2000 X-axis
0.15
–1600
0.1 –1200
0.05 –800
0

–2200 –2000 –1800 –1600 –1400 –1200 –1000 –800 –600


X-axis

(e) t=68 minutes 15 seconds

Fig. 8e—Fracture-geometry evolution for Example No. 1 case 1 (x- and y-coordinate in feet).

Even though this is a simple case, it illustrates some impor- and a net pressure of approximately 230 psi, while the width in NF
tant behaviors of complex fracture. As shown in Fig. 8a, a single 1 remains small because of a larger confining stress. The fractures
fracture is initially created at the perforation cluster. Because that branch off the ends of NF 1 also have a small width because
of the low fluid viscosity, the fracture width is relatively small, of low fluid viscosity and net pressure. The fracture branches that
approximately 0.05 in. When the fracture tips reach NF 1 and NF are directly opposing the main fracture have a smaller width com-
3, the friction coefficient and stress anisotropy are such that the pared with their opposite wings as a result of fracture-interference
hydraulic fracture is not able to cross the natural fractures. In the (stress-shadow) effect (Fig. 8d). These fractures then intersect NF 2
ensuing injection, the fracture length is not growing, but its width and NF 3 and are stopped by these natural fractures. As before, the
balloons up to 0.4 in. (Fig. 8b), while the net pressure increases pressure in these fracture branches starts to build up, and the width
from 40 to approximately 230 psi. balloons until the net pressure inside these fractures exceeds the
When the pressure is sufficiently large to exceed the maximum maximum horizontal stress to allow NF 2 to open up (Fig. 8e).
horizontal stress acting on the orthogonal NF 1, NF 1 is hydrauli- This example shows that when a hydraulic-fracture branch
cally opened. The fracture propagates quickly along NF 1 and intersects a natural fracture and cannot cross, the propagating
reaches the ends of the fracture, where the fracture branches off and branch is stopped and some other branches can continue to grow.
propagates in the preferred fracture direction again (Fig. 8c). The When all propagating tips are stopped by natural fractures, the net
orthogonal fracture NF 1 acts as a choke for the fluid in the main pressure inside the fracture network will have to rise to exceed the
fracture connected to the perforation and maintains a large width stress anisotropy. That is, the net pressure in the fracture system
is regulated by stress anisotropy. This can lead to much greater
1000 fracture width than predicted by the conventional biwing-fracture
1 2 model, or even the wiremesh model, in which straight fracture
planes is assumed. The net pressure response typically shows
a rapid rise in early time and then flattens after it exceeds the
maximum horizontal stress (Fig. 9). The high net pressure resulting
from width restriction and large pressure drop through the natural
Net Pressure Fracture 1 (psi)

fractures has also been observed by Jeffrey et al. (2009), though


the net pressure may not have to exceed the anisotropy to open up
the natural fracture in their 2D-plane-strain model.
Case 2: Friction Coefficient = 0.5, Stress Anisotropy = 500
100 psi, Intersection Angle = 90°. In this case, the stress anisotropy
is increased to 500 psi. At this condition, the crossing criterion
predicts that the hydraulic fracture will cross a natural fracture
when they intersect. Consequently, the UFM predicts a biwing
fracture as shown in Fig. 10. Similar geometry is obtained if the
stress anisotropy is kept at 200 psi, while the friction coefficient is
increased to greater than 0.7, which also corresponds to the condi-
tion of a hydraulic fracture crossing natural fractures.
Case 3: Friction Coefficient = 0.5, Stress Anisotropy = 500 psi,
10 Intersection Angle = 60°. In this case, the stress anisotropy is 500
0.1 1 10 100 psi, and the hydraulic fracture intersects natural fracture at an angle
Custom Pumping Time (min) of 60°. Fig. 11 shows the fracture geometry at t = 75 minutes, and
Fig. 12 shows the predicted net pressure. The figures show that with
Fig. 9—Net pressure for Example No. 1, Case 1. nonorthogonal intersection, the natural fractures stop the hydraulic

376 November 2011 SPE Production & Operations


X-axis
–2400 –2000 –1600 –1200 –800 –400

0.12
0.11
0.1
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0

–2400 –2000 –1600 –1200 –800 –400


X-axis

Fig. 10—Fracture geometry for Example No. 1, Case 2 (large anisotropy, t = 9 minutes, x- and y-coordinate in feet).

fractures from crossing, resulting in complex fracture rather than statistical parameters given in Table 3. Other input parameters are
biwing fracture as in Case 2 for an orthogonal intersection. Because the same as those given in Table 2. Stress anisotropy is assumed to
of the greater normal stress acting on the natural fractures (375 psi), be 200 psi. The pumping schedule is given in Table 4.
the fluid net pressure has to rise above 380 psi before the fluid pres- Fig. 13 shows the top view of the statistically generated natural
sure can open up NF 1. As a result, the fracture width in the main fractures and the predicted hydraulic-fracture network. Fig. 14
fracture is approximately 0.6 in., much wider than in the previous shows the fracture width contour and area proppant concentration
cases because of the greater net pressure. in 3D view. It shows that while a large hydraulic-fracture network
is created, proppant occupies only a small fraction of the fracture
Example No. 2: More-Complex Natural Fractures. In this exam- surface area created. This is mainly because of fast proppant
ple, more-complex natural fractures are generated on the basis of settling as a result of the low-viscosity fluid used, but it is also

X-axis Width
–2000 –1600 –1200 –800
0.7
0.65
0.6
0.55
0.5
0.45
0.4
0.35 –2400
0.3 –2000
0.25 X-axis
0.2 –1600
0.15 –1200
0.1
0.05 –800
0

–2000 –1600 –1200 –800


X-axis

Fig. 11—Fracture geometry for Example No. 1, Case 3 (large anisotropy, 60° angle, x- and y-coordinate in feet).

November 2011 SPE Production & Operations 377


Stage 1_Pumping schedule_1_Design1 More-detailed discussion of the impact of various formation
Net pressure (Fracture 1) parameters on the resulting fracture geometry and a field case
1000 study using the UFM are beyond the scope of this paper. Readers
1 2 are referred to the paper by Cipolla et al. (2011).

Conclusions
The new complex-hydraulic-fracture model presented in this paper
Net Pressure Fracture 1 (psi)

is capable of simulating propagation of hydraulic fractures in a


formation consisting of pre-existing natural fractures. Simulation
results from the model show that stress anisotropy, natural fractures,
100 and interfacial friction play critical roles in creating fracture net-
work complexity. Decreasing stress anisotropy or interfacial friction
can change the induced-fracture geometry from a biwing fracture
to a complex-fracture network for the same initial natural fractures.
The results illustrate the importance of rock fabrics and stresses on
fracture complexity in unconventional reservoirs. A complex-frac-
ture model provides a critical tool for engineers working on shale-
gas stimulation to be able to simulate the fracture geometry, prop-
10
pant distribution, and fracture conductivity; assess the influence
0.1 1 10 100 of natural fractures and various parameters on fracture geometry;
Custom Pumping Time (min) match the fracture geometry against the microseismic observations;
and optimize job design to maximize the production.
Fig. 12—Net pressure for Example No. 1, Case 3 (large anisot-
ropy, 60° angle).
Nomenclature
partially because of proppant bridging in the side fractures that Cnsij , Cnn
ij
, Cssij , Csnij = plane-strain elastic influence coefficients in 2D
have very small width. DD method
E' = plane-strain elastic modulus of the rock
H = fracture height
General Applications
Hfl = fluid height
As illustrated in the preceding simple examples, the pre-existing K' = fluid consistency index
natural fractures and the stress state in a formation (both minimum
KIc = rock toughness
and maximum horizontal stresses) have a major impact on the
hydraulic-fracture network created during a fracture treatment. L = fracture length
To predict the fracturing behavior, proper characterization of the n' = fluid power-law index
natural fractures is very important. However, detailed understand- P = absolute fluid pressure
ing of the natural fractures is often unavailable or incomplete. In pnet = net fluid pressure inside the fracture,
these circumstances, the UFM can be used to do “what-if” studies pnet (s, t) = p – σn
to assess the possible stimulation outcome for a range of uncer- q = flow rate of fluid
tain reservoir parameters. The predicted fracture geometry can be qL = leakoff rate per unit height through the fracture
compared to the effective stimulated volume as determined from surface into the rock
microseismic measurement, and these measurements can be used Q = pumping rate
to help constrain the uncertain parameters, leading to a more-pre- t = time
dictive design model. Furthermore, the predicted fracture geometry
T0 = tensile strength of rock
needs to be integrated with a production simulator to assess the
potential production from the fracture stimulation, which allows Pf = unconventional-fracture-model fluid density
the production engineer to match the production performance σn = horizontal stress acting perpendicular to fracture
further and to evaluate design changes to optimize the production. element
All these techniques require close integration of a fracture model σs = shear stress acting on fracture element
such as the UFM with reservoir characterization, geomechanical w = average fracture opening at vertical cross section
modeling, microseismic interpretation, and production-simulation
tools. The UFM has been integrated into a software platform that Acknowledgment
enables these integrations. The general workflow can be found in The authors wish to thank Schlumberger for permission to publish
Cipolla et al. (2010). this paper.

TABLE 3—NATURAL-FRACTURE PROPERTIES FOR EXAMPLE NO. 2

Length (ft) Orientation Spacing (ft) Friction Coeff. Cohesion (psi)

Average 200 120° 100 0.5 0


Std. deviation 50 10° 50 0 0

TABLE 4—PUMP SCHEDULE

Stage Name Pump Rate (bpm) Fluid Fluid Vol (gal) Proppant Prop Conc. (ppga) Proppant Mass (lbs)

Pad 50 Slickwater 100,000 40/70 mesh sand 0 0


0.5 ppga 50 Slickwater 100,000 40/70 mesh sand 0.5 50,000
0.75 ppga 50 Slickwater 100,000 40/70 mesh sand 0.75 75,000
1 ppga 50 Slickwater 100,000 40/70 mesh sand 1.0 100,000

378 November 2011 SPE Production & Operations


X-axis
–3200 –2800 –2400 –2000 –1600 –1200 –800 –400 0 400

2000

1800

1600

1400

1200

1000

800

600

400

200

–200

–400

–600

–800

–3200 –2800 –2400 –2000 –1600 –1200 –800 –400 0 400


X-axis

Fig. 13—Natural fractures and created-hydraulic-fracture geometry for Example No. 2 (x- and y-coordinate in feet).

References Cipolla, C.L., Weng, X., Mack, M.G., et al. 2011. Integrating Microseismic
Adachi, J., Seibrits, E., Peirce, A., and Desroches, J. 2007. Computer simulation Mapping and Complex Fracture Modeling to Characterize Hydraulic
of hydraulic fractures. Int. J. Rock Mech. Min. Sci. & Geomech. Abstracts Fracture Complexity. Paper SPE 140185 presented at the SPE Hydrau-
44 (5): 739–757. http://dx.doi.org/10.1016/j.ijrmms.2006.11.006. lic Fracturing Technology Conference, The Woodlands, Texas, USA,
Blanton, T.L. 1982. An Experimental Study of Interaction Between Hydrau- 24–26 January. http://dx.doi.org/10.2118/140185-MS.
lically Induced and Pre-Existing Fractures. Paper SPE 10847 presented Cipolla, C.L., Williams, M.J., Weng, X., Mack, M.G., and Maxwell,
at the SPE Unconventional Gas Recovery Symposium, Pittsburgh, S.C. 2010. Hydraulic Fracture Monitoring to Reservoir Simulation:
Pennsylvania, USA, 16–18 May. http://dx.doi.org/10.2118/10847-MS. Maximizing Value. Paper SPE 133877 presented at the SPE Annual

Width Proppant–Areal_Distribution
0.5 5.5
0.45 5
0.4 4.5
0.35 400 ft 4
3.5
0.3
0.25 3
0.2 2.5
0.15 2
1.5
0.1 1
0.05 0.5
0 0

Fig. 14—Fracture width and proppant area concentration for Example No. 2.

November 2011 SPE Production & Operations 379


Technical Conference and Exhibition, Florence, Italy, 19–22 Septem- presented at the 42nd US Rock Mechanics Symposium, San Francisco,
ber. http://dx.doi.org/10.2118/133877-MS. 29 June–2 July.
Crouch, S.L. and Starfield, A.M. 1983. Boundary Element Methods in Renshaw, C.E. and Pollard, D.D. 1995. An experimentally verified criterion
Solid Mechanics: With Applications in Rock Mechanics and Geological for propagation across unbounded frictional interfaces in brittle, linear
Engineering. London: George Allen & Unwin. elastic materials. Int. J. Rock Mech. Min. Sci. & Geomech. Abstracts
Daneshy, A.A. 1978. Numerical Solution of Sand Transport in Hydraulic 32 (3): 237–249. http://dx.doi.org/10.1016/0148-9062(94)00037-4.
Fracturing. J Pet Technol 30 (1): 132–140. SPE-5636-PA. http://dx.doi. Schiller, L. and Naumann, A. 1933. A drag coefficient correlation. VDI
org/10.2118/5636-PA. Zeitschrift 77: 318–320.
Daniels, J., Waters, G., Le Calvez, J., Lassek, J., and Bentley, D. 2007. Thiercelin, M. and Makkhyu, E. 2007. Stress field in the vicinity of a natu-
Contacting More of the Barnett Shale Through an Integration of Real- ral fault activated by the propagation of an induced hydraulic fracture.
Time Microseismic Monitoring, Petrophysics, and Hydraulic Fracture Proc., 1st Canada-US Rock Mechanics Symposium (Rock Mechanics:
Design. Paper SPE 110562 presented at the SPE Annual Technical Con- Meeting Society’s Challenges and Demands), Vancouver, Canada,
ference and Exhibition, Anaheim, California, USA, 11–14 November. 27–31 May, 1617–1624.
http://dx.doi.org/10.2118/110562-MS. Wang, J., Joseph, D.D., Patankar, N.A., Conway, M., and Barree, R.D.
Fisher, M.K., Wright, C.A., Davidson, B.M., et al. 2002. Integrating 2003. Bi-power law correlations for sediment transport in pressure
Fracture Mapping Technologies to Optimize Stimulations in the Bar- driven channel flows. Int. J. Multiphase Flow 29 (3): 475–494. http://
nett Shale. Paper SPE 77441 presented at the SPE Annual Technical dx.doi.org/10.1016/s0301-9322(02)00152-0.
Conference and Exhibition, San Antonio, Texas, USA, 29 September– Warpinski, N.R. and Teufel, L.W. 1987. Influence of Geologic Discontinui-
2 October. http://dx.doi.org/10.2118/77441-MS. ties on Hydraulic Fracture Propagation. J Pet Technol 39 (2): 209–220.
Gu, H. and Weng, X. 2010. Criterion For Fractures Crossing Frictional SPE-13224-PA. http://dx.doi.org/10.2118/13224-PA.
Interfaces At Non-orthogonal Angles. Paper ARMA 10-198 presented Xu, W., Calvez, J.H.L., and Thiercelin, M.J. 2009. Characterization of
at the 44th U.S. Rock Mechanics Symposium and 5th U.S.-Canada Hydraulically-Induced Fracture Network Using Treatment and Micro-
Rock Mechanics Symposium, Salt Lake City, Utah, USA, 27–30 seismic Data in a Tight-Gas Sand Formation: A Geomechanical
June. Approach. Paper SPE 125237 presented at the SPE Tight Gas Comple-
Gu, H., Weng, X., Lund, J.B., Mack, M.G., Ganguly, U., and Suarez- tions Conference, San Antonio, Texas, USA, 15–17 June. http://dx.doi.
Rivera, R. 2011. Hydraulic Fracture Crossing Natural Fracture at Non- org/10.2118/125237-MS.
Orthogonal Angles, A Criterion, Its Validation and Applications. Paper Xu, W., Thiercelin, M.J., Ganguly, U., et al. 2010. Wiremesh: A Novel
SPE 139984 presented at the SPE Hydraulic Fracturing Technology Shale Fracturing Simulator. Paper SPE 132218 presented at the Inter-
Conference, The Woodlands, Texas, USA, 24–26 January. http://dx.doi. national Oil and Gas Conference and Exhibition in China, Beijing, 8–10
org/10.2118/139984-MS. June. http://dx.doi.org/10.2118/132218-MS.
Gulrajani, S.N., Nolte, K.G., and Romero, J. 1997. Evaluation of the M-Site Zhang, X. and Jeffrey, R.G. 2006. The role of friction and secondary flaws
B-Sand Fracture Experiments: The Evolution of a Pressure Analysis on deflection and reinitiation of hydraulic fractures at orthogonal pre-
Methodology. Paper SPE 38575 presented at the SPE Annual Technical existing fractures. Geophys. J. Int. 166 (3): 1454–1465. http://dx.doi.
Conference and Exhibition, San Antonio, Texas, USA, 5–8 October. org/10.1111/j.1365-246X.2006.03062.x.
http://dx.doi.org/10.2118/38575-MS. Zhang, X., Jeffrey, R.G., and Thiercelin, M. 2007. Effects of Frictional
Jeffrey, R.G., Zhang, X., and Thiercelin, M. 2009. Hydraulic Fracture Geological Discontinuities on Hydraulic Fracture Propagation. Paper
Offsetting in Naturally Fractured Reservoirs: Quantifying a Long- SPE 106111 presented at the SPE Hydraulic Fracturing Technology
Recognized Process. Paper SPE 119351 presented at the SPE Hydrau- Conference, College Station, Texas, USA, 29–31 January. http://dx.doi.
lic Fracturing Technology Conference, The Woodlands, Texas, USA, org/10.2118/106111-MS.
19–21 January. http://dx.doi.org/10.2118/119351-MS.
Le Calvez, J.H., Klem, R.C., Bennett, L., Erwemi, A., Craven, M., and Pala- Xiaowei Weng is a principal engineer and project manager
cio, J.C. 2007. Real-Time Monitoring of Hydraulic Fracture Treatment: of modeling and mechanics with Schlumberger in Sugar
A Tool to Improve Completion and Reservoir Management. Paper Land, Texas, USA. His research interests include hydraulic-frac-
SPE 106159 presented at the SPE Hydraulic Fracturing Technology ture modeling, acid fracturing, multifractured horizontal well
Conference, College Station, Texas, USA, 29–31 January. http://dx.doi. completion and production, wellbore hydraulics, and coiled-
org/10.2118/106159-MS. tubing cleanout. He received holds MS and PhD degrees in
LeVeque, R.J. 1992. Numerical Methods for Conservation Laws, second engineering mechanics from The University of Texas at Austin.
edition, 165. Basel, Switzerland: Lectures in Mathematics, Birkhaüser He is a member of SPE. Olga Kresse is a senior engineer with
Verlag. Schlumberger in Sugar Land, Texas, USA. Her primary research
interests are fracture and fluid mechanics, acoustics, and
Mack, M.G. and Warpinski, N.R. 2000. Mechanics of Hydraulic Fracturing.
applied math. Her most recent work is concentrated on
In Reservoir Stimulation, third edition, ed. M.J. Economides and K.G. numerical modeling of hydraulic fracturing in unconventional
Nolte, Chap. 6. West Sussex, UK: John Wiley & Sons. reservoirs. She holds two PhD degrees – one in solid mechanics
Maxwell, S.C., Urbancik, T.I., Steinsberger, N.P., and Zinno, R. 2002. from T. Shevchenko National University and the other in fracture
Microseismic Imaging of Hydraulic Fracture Complexity in the Bar- mechanics from the University of Minnesota. She is a member
nett Shale. Paper SPE 77440 presented at the SPE Annual Technology of SPE, ASME, ARMA, AGU, and ASA. Charles-Edouard Cohen is
Conference and Exhibition, San Antonio, Texas, USA, 29 September– a modeling engineer with Schlumberger in Sugar Land, Texas,
2 October. http://dx.doi.org/10.2118/77440-MS. USA. His research interests include modeling and simulation,
Meyer, B.R. and Bazan, L.W. 2011. A Discrete Fracture Network Model fluid mechanics, matrix acidizing in carbonate reservoir, frac-
turing, and production. His most recent works are on numeri-
for Hydraulically Induced Fractures - Theory, Parametric and Case
cal modeling of hydraulic fracturing in unconventional reser-
Studies. Paper SPE 140514 presented at the SPE Hydraulic Fracturing voirs. He holds a PhD degree in reservoir engineering from the
Technology Conference, The Woodlands, Texas, USA, 24–26 January. French Institute for Petroleum and the National Polytechnique
http://dx.doi.org/10.2118/140514-MS. Institute of Toulouse, an MSc degree in hydraulics and fluid
Nolte, K.G. 1991. Fracturing-Pressure Analysis for Nonideal Behav- mechanics, and an MSc degree in fluid dynamics from the
ior. J Pet Technol 43 (2): 210–218. SPE-20704-PA. http://dx.doi.org/ ENSEEIHT in Toulouse, France. He is a member of SPE and ARMA.
10.2118/20704-PA. Ruiting Wu is a geomechanics engineer with Schlumberger.
Olson, J.E. 2004. Predicting fracture swarms—the influence of subcritical Her research interests include fracture mechanics and sand
crack growth and the crack-tip process zone on joint spacing in rock. In management. She holds a PhD degree from the Georgia
Institute of Technology. She is a member of SPE. Hongren Gu
The Initiation, Propagation, and Arrest of Joints and Other Fractures,
is a principal engineer with Schlumberger in Sugar Land, Texas,
ed. J.W. Cosgrove and T. Engelder, No. 231, 73–87. Bath, UK: Special USA. His research interests include reservoir stimulation, fracture
Publication, Geological Society Publishing House. mechanics, and wellbore hydraulics. He holds an MS degree
Olson, J.E. 2008. Multi-fracture Propagation Modeling: Application to from Xian Jiaotong University in China and a PhD degree from
hydraulic fracturing in shales and tight gas sands. Paper ARMA 08-327 University of Texas at Austin, both in engineering mechanics.

380 November 2011 SPE Production & Operations

Вам также может понравиться