Sunteți pe pagina 1din 28

SPE 165681

Evaluation of Quintuple Porosity in Shale Petroleum Reservoirs


Bruno Lopez and Roberto Aguilera, Schulich School of Engineering, University of Calgary

Copyright 2013, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Eastern Regional Meeting held in Pittsburgh, Pennsylvania, USA, 2022 August 2013.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract

The use of a quintuple porosity system for calculation of original petroleum in place (OPIP) in shales is important as
neglecting some of the porosities can result in pessimistic values of OPIP and production rates. Based on the concept of Total
Petroleum System (Magoon and Beaumont, 1999) the word Petroleum includes (1) thermal and biological hydrocarbon gas,
(2) condensates, (3) crude oils and (4) natural bitumen. In the case of natural gas, the gas is trapped and stored in shale in
different ways: (1) gas adsorbed in the kerogen material, (2) free gas trapped in nonorganic inter-particle (matrix) porosity,
(3) free gas trapped in microfracture and slot porosity, (4) free gas stored in hydraulic fractures created during the stimulation
of the shale reservoir, and (5) free gas trapped in a pore network developed within the organic matter or kerogen. An
additional storage element is provided by gas dissolved in kerogen.

The governing equations that describe the gas mass balance in the quintuple porosity model are presented in detail. The series
and parallel gas transport approaches discussed previously in the literature are shown to be special cases of the new general
gas transport formulation developed in this study. The effects of permeability stress-dependency are taken into account. Real
data from Devonian gas shales are used to illustrate the effect of free gas, adsorbed gas and dissolved gas in a material
balance crossplot of P/Z vs. cumulative gas production.

Historically, the large contribution of organic porosity, natural fractures and hydraulic fractures that can contribute a
significant amount of free petroleum in place has not been taken into account. And many of the laboratory experiments used
for determining data utilized in computations of OPIP and production rates have been carried out in crushed samples, which
by their very nature do not generally preserve natural fractures, slots and all the porosity present in the organic matter. This
leads to pessimistic values of OPIP and rates. This helps to explain the larger than anticipated rates and recoveries of natural
gas from some of these formations, for example Devonian shales, which have been producing for several decades.

Although the quintuple porosity characterization mentioned above indicates very heterogeneous systems, the production
performance is less heterogeneous than that of carbonates, sandstones and naturally fractured tight reservoirs. This surprising
result is demonstrated with the use of actual production data from various petroleum reservoirs around the world.
The subject matter is significant because of the large volume of petroleum resources in shales throughout the world, which
probably are underestimated because of not considering in a single model all types of porosity discussed in this study.

Introduction

Based on the concept of Total Petroleum System (Magoon and Beaumont, 1999) the word Petroleum includes (1) thermal
and biological hydrocarbon gas, (2) condensates, (3) crude oils and (4) natural bitumen. These hydrocarbons and rocks form
part of the Total Petroleum System which includes all the geologic elements and processes required for an oil and gas
accumulation to exist. The Petroleum System is a unifying concept that encompasses all of the disparate elements and
processes of petroleum geology including a pod of active source rock and all genetically related oil and gas accumulations.
(Magoon and Beaumont, 1999). The concept has been used by Aguilera (2013) to explain flow units as a continuum that
occurs between conventional, tight gas, shale gas, tight oil and shale oil reservoirs. The word petroleum as used in current
study concentrates particularly on natural gas stored in shales.
2 SPE 165681

Shale petroleum reservoirs are found throughout the world. A problem is that there is significant uncertainty regarding the
amount of petroleum that can be recovered from shale plays. In order to reduce such uncertainty, it is important to have a
better understanding of the physics, types of storage and transport mechanisms that control fluid flow through the nanoscale
media in shale reservoirs. Previous work along these lines include petrophysical descriptions (Aguilera, 2010), analysis of
non-Darcy flow mechanisms in nanoscale porous media (Javadpour, 2009; Civan, 2010), and development of numerical
simulation approaches (Akklutu and Fathi, 2011; Hudson, 2011; Shabro et al., 2011, Andrade et al., 2011; Hudson et al.,
2012; Shabro et al., 2012; Swami, 2013). This paper presents new mathematical models for incorporating the dominant
storage and flow mechanisms that occur in shales.

But first, some geologic aspects: Generally, economic shale petroleum reservoirs are organic rich and are characterized by
different types of pores in the organic and non-organic (mineral) matter. Geologically there are different types of pores in
shales including interparticle, intraparticle and fracture pores as shown in Figure 1 (Loucks et al., 2012, Cheadle and Jiang,
2013).

The interparticle porosity can be found between grains and/or crystals in the mineral matter. Intraparticle porosity can found
in the organic as well as the non-organic matter. The presence of porous networks in the organic material of shales was first
brought to the petroleum engineering literatures attention by Wang and Reed (2009). Wang and Reed also discussed the
presence of four different types of porosity: mineral matrix, fracture, organic and hydraulic. Intraparticle porosity in the
organic matter has more probabilities of being effective as compared with the intraparticle porosity in the non-organic matter
because of the interconnectivity of organic matter particles (Loucks et al., 2012).

Naturally fractured pores and their morphology play an important role in shale petroleum reservoirs. Natural fractures can be
completely or partially cemented. Handwerger et al. (2012) have indicated that observation on thousands of feet of cores
show a pervasive presence of subvertical mineralized fractures (veins) on most vertical cores, with few exceptions. In some
cases such as the Haynesville Shale, the mineralized veins are oriented subhorizontally, suggesting the presence of veins is
pervasive and represents an important property of these systems. These types of cemented fractures play a significant role in
stimulation of shale gas and shale oil reservoirs.

On the other hand of the spectrum there are natural fractures that can enhance permeability significantly. This is the case of
the Bakken in North Dakota (Elm Coulee, Parshall, and Sanish pools) where the host permeability to oil can be enhanced by
natural fractures stemming from dolomitization (Sonnenberg, 2011) and slot porosity. And to add complications to the
characterization shale problem there are instances where researchers have not been able to observe abundant open
microfractures at the thin-section or SEM scale even after studying hundreds of samples. However, these researchers
emphasize the importance of natural microfractures as indicated in Figure 1 (Loucks et al., 2012).

The clear message then is that shale reservoirs are different and each one has to be considered as a research problem by itself
to improve the probabilities of success. What works nicely in a given shale reservoir might fail miserably in the next one.

The qualitative geological descriptions summarized briefly above are critical for quantitative engineering assessment of shale
petroleum reservoirs. For our engineering purposes we call the interparticle porosity in the non-organic (mineral) matter
discussed above matrix porosity m, the natural fracture porosity 2, and the intraparticle organic porosity SEM.

The assessment proposed in this paper requires the introduction of porosity created during the stimulation of the shale
petroleum reservoir (hf). Furthermore, in the case of natural gas reservoirs the intraparticle organic porosity SEM is divided
into the organic porosity within which gas is adsorbed (ads_c) and the organic porosity within which gas is free (org). The
summation of m + 2 + org + ads_c + hf made-up the quintuple porosity system considered in this study. In addition, the
paper also considers an additional storage mechanism provided by gas dissolved in kerogen.

Porosity and Permeability Review

To try to handle the above types of porosity quantitatively as well as the porosity created during stimulation jobs Aguilera
(2010) indicated that shale gas reservoirs were more realistically represented by at least a quadruple porosity model
represented by the equation,

!! = ! + !"# + ! + !! (1)
SPE 165681 3

where,

m = nonorganic matrix porosity scaled to the bulk volume of the composite system, fraction.
org = organic matrix porosity scaled to the bulk volume of the composite system, fraction.
2 = natural fracture porosity scaled to the bulk volume of the composite system, fraction.
hf = hydraulic fracture porosity scaled to the bulk volume of the composite system, fraction.

In this paper an additional porosity (ads_c) is defined leading to a quintuple porosity system. The nonorganic matrix porosity
is related to the inter-particle porosity associated with the presence of clays, silica and other minerals. Organic matrix
porosity refers to the pores identified within the kerogen as observed in Figure 2. Most shale gas reservoirs are naturally
fractured. If the fractures are opened or partially cemented they provide some additional storage and permeability. If they are
completely cemented they affect hydraulic fracturing growth in a positive way.

In their natural state, shale gas reservoirs, even with natural microfractures, exhibit very low permeability. Therefore, they
have to be hydraulically fractured to obtain commercial production rates. As a result, the contribution of hydraulic fractures is
taken into account for calculation of total porosity. Aguilera (2010) argues that at least a quadruple porosity formulation is
more accurate than a triple porosity system (without hydraulic fractures contribution).

Subsequently, Andrade et al. (2011) also indicated that shales could be represented by a multiple-porosity system comprised
by organic matter, inorganic matter, natural fractures and induced fractures. Based on this quad-media continuum approach,
Civan (2012) defined six exchange pathways for connectivity between the different components of shales as shown in Figure
3. Civan (2012) suggested the following sequence of gas transport out of shale: (1) free gas, (2) adsorbed gas and (3)
dissolved gas within organic matter and water.

Due to the presence of pores with sizes in the range of nanometers, non-Darcy flow mechanisms can take place in some parts
of the shale reservoirs. These different flow conditions are described by the Knudsen number limits shown in Table 1. Civan
(2010) considered non-Darcy mechanisms in the definition of an apparent permeability expression as a function of Knudsen
number. The apparent permeability corresponding to gaseous viscous flow, slip flow and Knudsen diffusion is as follows,

! !!! ! !!!
! !! !!! ! !"
! = + 1 (2)
!!! !!! !" !!! !" ! !"

where F, the wall slip correction factor (Brown et al., 1946),

! ! !!! !
=1+4 1 (3)
! !! ! !!

and other symbols are explained in the nomenclature.

Some researchers (Cui et al., 2009; Ambrose et al., 2010) have tried to explain the amount of gas adsorbed on kerogen pore
walls with the use of a sorbed porosity. Cui et al. (2009) suggested that adsorption represents an additional gas storage
mechanism that increases the total gas capacity of shales. This incremental gas capacity was introduced in the accumulation
term of the mass balance equation, and adding the assumption that gas desorption can be approximated by the Langmuir
isotherm, led to the following equation for calculating effective adsorption porosity,

!! !! !! !! !!
! = (4)
!!"# !! ! !! !! !

Note that a is not based on the usual concept of porosity, which relates void space to the bulk volume of the rock. Ambrose
et al. (2010) proposed a methodology to perform original gas in place (OGIP) calculations, which are affected by the volume
occupied by adsorbed gas. The modification of volumetric equations utilized to calculate total and free gas capacity led them
to establish a correction factor defined as sorbed phase porosity fraction,

!! !
! = 1.31810!! !" (5)
!! !! !!

Ambrose et al. (2010) do not present a detailed derivation of the above expression but as in the case of equation 4 it does not
appear to be related to the usual concept of porosity. However, they mention that the fractional volume occupied by the
sorbed phase can be calculated based on the density of the adsorbed phase, bulk density of the rock, and molecular weight of
the adsorbed phase.
4 SPE 165681

In this paper we define a real adsorbed porosity. We call it real since our adsorbed porosity is established on the basis of the
usual definition of porosity, i.e., the ratio of pore volume to bulk volume of a rock. The reason for developing a real adsorbed
porosity is that in this way, the gas stored as an adsorbed phase can be included into the total porosity quantification and
volumetric quantification of OGIP in shale reservoirs. Furthermore, the adsorbed real porosity is a key component of the
mass balance equation for organic matrix we are proposing in this paper. This porosity can change or remain approximately
constant once the reservoir goes on production. This highlights the importance of the quintuple porosity system definition for
shales introduced in this paper. Intrinsic and scaled to bulk volume adsorbed porosities are presented and discussed later in
this paper.

Unfortunately, there is not a well-defined protocol for estimating porosity storage and permeability in shale reservoirs.
Handwerger et al. (2012) indicate their preference for running their experiments on crushed rock for tight shale
applications. They argue that core from shales provide optimistic results because of significant induced microfracturing
while coring stemming from colloidal size of grains, planar shape grains, mixture of organic and inorganic components and
weak contacts.

On the other hand Sigal et al. (2013) indicate their preference for gas storage measurements at reservoir conditions on a
solid core sample. They argue that current methods that use crushed samples to measure adsorption with reservoir gas but
pore volume with helium do not provide accurate gas storage results.

Cores of shales that produce oil present additional problems related to the increased liquid hydrocarbon content and phase
changes due to variations in pressure and temperature (Handwerger et al., 2012). To stay away from PVT complications
(subject of another study) the bulk of the development presented in this paper is carried out with the use of natural gas. The
governing equations that describe the gas mass balance in each porous medium are presented in detail later in this paper.

Permeabilities from cores and drill cuttings are very small. The same holds true for permeabilities calculated from production
data. For example, Tran et al. (2011) have reported a permeability of 0.013 md while working with production characteristics
of the Bakken, one of the most prolific shale oil reservoirs. Although porosities and permeabilities from shale cores and
cuttings are very small, they can be used for establishing flow units as shown by Aguilera (2013).

Current Shale Gas Formulations

Over the past few years, several studies have been undertaken to try to gain a better understanding of gas transport within
organic nano porous media. They all provide innovative alternatives to conventional methods. The main differences between
these models are related to components considered in the total shale system, the connectivity between each component, and
the ways in which gas is trapped in shales. Each one of these studies is summarized and critiqued briefly in this section with a
view to identify similarities and differences between the existing formulations and the quintuple porosity model proposed in
this work.

Akkutlu and Fathi (2011)

Akkutlu and Fathi (2011) developed a mathematical model to analyze multi-scale gas transport phenomena in organic-rich
shale core plugs. This formulation takes into account both organic and inorganic matrix and also natural fractures. Diffusion
and nonlinear sorption processes are considered within the organic matter. They provide a numerical solution using a finite
difference method to discretize each of the mass balance equations proposed.

The gas transport in organic-rich shales follows a series approach, in which the organic matrix (kerogen) communicates with
the inorganic matrix, and this one releases gas to the fracture medium. Figure 4 shows a schematic of the model used by
Akkutlu and Fathi (2011). Details of the derivation of the governing equations for gas storage and transport can be found in
their original publication.

This formulation by itself is robust and represents an interesting approach to be applied in the development of a new-
generation of shale simulators. Nevertheless, the assumption of gas transport in series can be modified in order to include
additional interaction mechanisms between the various media. Also, the storage within hydraulic fractures should be taken
into account since most shale reservoirs are hydraulically fractured.

Shabro et al. (2011)

Shabro et al. (2011) described the production history and pressure distribution in shale reservoirs based on a combination of a
pore-scale model with a reservoir-scale model. The gas transport in porous media includes the effects of no-slip and slip
flow, Knudsen diffusion, and Langmuir desorption. The adsorption/desorption process is analyzed under non-equilibrium
SPE 165681 5

conditions, which are handled with the definition of a surface mass balance equation. The contributions of the non-Darcy
flow regimes to the total gas flow were assessed by the use of the apparent permeability definition proposed by Javadpour
(2009).

According to their results, Langmuir desorption from organic nanopores is an important aspect in the quantification of the gas
in place in shale formations. Gas desorption acts as a pressure support to increase the production obtained from organic
shales. Figure 5 shows the incremental in the production due to the effect of gas desorption.

Although this formulation is innovative and explains the observed higher than predicted production in shales, there are some
assumptions that could be changed to improve the modeling of these reservoirs. The gas compressibility factor was assumed
to be constant; therefore, it was not included in the analysis. In addition, governing equations were solved explicitly rather
than finding the solution via an implicit formulation. Also, the proposed model considers the organic matrix as the total
system; whereas inorganic matrix, natural fractures and hydraulic fractures are neglected.

Hudson (2011)

Hudson (2011) analyzed the gas transport in shales with an explicit differential model. In this analysis, shale reservoirs are
represented by the quadruple porosity system introduced by Aguilera (2010) and Andrade et al. (2011): organic and inorganic
matrix, and natural and hydraulic fractures. Langmuir desorption for the organic porosity system is also considered in this
study. The connection between matrices is established under both series and parallel conditions. Figure 6 shows how the gas
transport for the parallel case was defined. It can be seen from Figure 6 that there is not any interaction between organic and
inorganic matrices.

The simulation of this quad-porosity model is done with the use of a commercial simulator. The differential equations
derived from a conservation approach were solved using the finite difference approximation included in the commercial
simulator.

The methodology presented by Hudson (2011) represents an interesting practical approach to simulate shale reservoirs due to
the lack of proper shale simulators. However, there are some changes that could be considered to improve the results
obtained from the simulation. For example, connectivity between organic and inorganic matrices and gas diffusion from
kerogen are two physical phenomena that should be considered in the analysis of organic shales.

Shabro et al. (2012)

Shabro et al. (2012) studied the effects of gas diffusion in kerogen, Langmuir desorption, slip flow and Knudsen diffusion on
the gas production of shale reservoirs. They used the same approach as in their previous work (Shabro et al., 2011) but added
a diffusion equation to analyze gas flow in kerogen.

The impact of each flow mechanism on total gas production was evaluated by comparing several scenarios. They found that
gas diffusion increases gas production. The gas that diffuses from kerogen feeds the surfaces of the organic pores letting the
adsorption of molecules of gas to the pores walls to continue. The continuous adsorption keeps the desorption process going
as reservoir pressure decreases. The contribution of gas diffusion is shown in Figure 7.

Despite the valid fundamental physics considered in the development of their reservoir model, the authors cite limitations
associated with a key assumption made for their development: neglecting the effect of fractures.

Swami (2013)

Swami (2013) provided numerical solution to both pore scale and laboratory scale models. He developed theoretical models
for a quad porosity system in shale reservoirs. Quad porosity in this model does not refer to the presence of four porosity
components but to four storage mechanisms: (1) free gas in natural fractures, (2) free gas in organic matrix nanopores, (3)
adsorbed gas and (4) gas dissolved in kerogen. Strictly speaking, the model only considers two porosities: natural fractures
and organic matrix. Non-Darcy flow mechanisms (Knudsen diffusion and slip flow) were also taken into account in the form
of an apparent permeability expression (Swami and Settari, 2012), which results from the derivation of the pore scale model
after consideration of slip flow and Knudsen diffusion. The apparent permeability term is expressed as,

! = ! ! + ! (6)
6 SPE 165681

where km and Dm are respectively expressed as,

!
! = ! (7)
!

!
! = (8)
!

All the symbols are explained in the nomenclature. The above apparent permeability is the same as the one proposed by
Civan (2010). After some arrangements in Eq. 2, the expression for apparent permeability (Eq. 6) is obtained as shown in
Appendix A.

Swami (2013) concluded that cumulative production of gas from shale reservoirs is really sensitive to the presence of the
mechanisms included in his study. He highlights the importance of the inclusion of the aforementioned mechanisms for the
modeling and production forecasting of shale gas reservoirs. Figure 8 shows the effects of each source of gas on the total
production. These effects are corroborated against experimental data published by Javadpour et al. (2007). The solid lines
represent the results from the numerical solution developed by Swami (2013) whereas the points are the experimental data
obtained by Javadpour et al. (2007). Figure 8 shows an excellent agreement between simulated and laboratory data.

Swami (2013) draws attention to improvements that can be done to his model. He recommends considering the effects of the
inorganic matter since this could affect the results and interpretation. The gas transport in series could be modified in order to
accommodate the inclusion of inorganic matter. In addition, the contribution of hydraulic fractures should be analyzed and a
transient interaction between the different media should be applied instead of a pseudo steady state formulation as suggested
by Brown et al. (2011).

Quintuple Porosity Formulation

Shale matrix is composed of inorganic matter (clay, silica and other minerals) and organic matter (kerogen). Many shale gas
reservoirs are characterized by natural fractures and low matrix permeability values. This fact imposes the need for multi-
stage hydraulic fracturing of horizontal wells to make their production economically viable. Furthermore, gas adsorption to
the internal surface of the organic pores represents an important storage mechanism of gas in shale reservoirs. Gas diffusion
from kerogen (Javadpour et al., 2007; Akkutlu and Fathi, 2011; Swami and Settari, 2012; Shabro et al., 2012; Swami, 2013;
Moghanloo et al., 2013) is an important source of gas to be added for quantification of total gas in place in shales. The result
is a composite system in shale reservoirs made up of dissolved gas in kerogen plus the following storage mechanisms,

1. Adsorbed gas in the kerogen material


2. Free gas trapped in nonorganic inter-particle (matrix) porosity
3. Free gas trapped in microfracture and slot porosity
4. Free gas stored in hydraulic fractures created during stimulation of the shale reservoir
5. Free gas trapped in an intra-particle pore network developed within the organic matter or kerogen

The aforementioned storage mechanisms form the quintuple porosity system for shale gas reservoirs is introduced in this
paper. In the case of oil, it can be trapped in the inter-particle porosity, natural fracture porosity and the stimulated reservoir
volume. Figure 9 suggests that every porosity component of the system can be connected to each other, which is physically
possible in complex systems such as shale gas reservoirs. However, the complete physical phenomenon at the reservoir scale
cannot be handled mathematically because its representation would require a huge number of equations. The solution of this
problem would require the use of considerable and prohibitive computer power and a significant amount of time.

Because of this, from a practical point of view, gas production from shales can be represented by the schematics shown in
Figures 10 and 11. In Figure 10 organic porosity is connected to inorganic porosity. Both of them feed natural fractures,
which contribute to gas production in two ways: (1) direct communication with the production well, and (2) feeding gas to
hydraulic fractures, which in turn are connected to the production well.

Note that the formulation shown in Figure 10 simplifies to the gas transport in series and parallel discussed by other authors
(Akklutu and Fathi, 2011; Hudson, 2011; Shabro et al., 2011; Hudson et al., 2012; Shabro et al., 2012; Swami, 2013) as
shown by the red solid lines in Figure 11. Therefore, the series and parallel approaches discussed previously in the literature
are special cases of the general gas transport formulation developed in this paper.

The mathematical accommodation of the adsorbed gas in the quadruple porosity system defined by Aguilera (2010) is
presented next. This inclusion yields the quintuple porosity system introduced in this study.
SPE 165681 7

Real adsorbed porosity

Assuming a circular geometry of the pores within the organic matter (Sakhaee-Pour and Bryant, 2012) based on SEM
images, a thickness of the adsorbed layer (Ambrose et al., 2010; Sakhaee-Pour and Bryant, 2012) can be identified. A
schematic of this situation is shown in Figure 12. The thickness of the adsorbed layer occupies a volume within the organic
pore. This volume can be visualized as pore volume since it is a space available to store gas. The adsorbed pore volume can
be related to a the bulk volume of only the kerogen or the bulk volume of the whole rock in such a way that a real adsorbed
porosity (ads) can be computed for example from,

!!_!"#
!"# = (9)
!!

The thickness of the adsorbed layer (hads) can be calculated with the use of two different methods: (1) an approach used for
calculation of protein film thickness (Vasina et al., 2009), and (2) characterization of the structure of adsorption layers
formed by adsorption of polymers (Chornaya et al., 2002). Appendix B shows detailed example calculations of the thickness
of the adsorbed layer using the aforementioned methods. According to the first method, for a temperature of 21 C and a
pressure of 28 MPa, the thickness of the methane adsorbed layer is approximately equal to 0.63 nm. This value is consistent
with the average total thickness for an adsorbed methane layer suggested by Ambrose et al. (2010). In this regard, it must be
noted that Ambrose et al. (2010) suggest a typical value of 0.7 nm for the same pressure and temperature. However, they do
not show how their value was calculated.

The second method calculates thickness of the adsorbed layer by assuming the surface area of a cylindrical pore. In this case,
the calculated thickness of methane adsorbed layer is equal to 0.39 nm as shown in Appendix B. There are various methods
in the literature for calculation of the surface area available for adsorption (Brooks and Purcell, 1952; Perez Rosales, 1967;
Kumar and Fatt, 1970; Donaldson et al., 1975; Barlai, 1976; Mortensen et al., 1998; Basbug and Karpyn, 2007).

If a circular cross sectional area of the organic pore is assumed as in Figure 12, a cylindrical geometry can be considered to
estimate the volume associated with the thickness of the adsorbed layer. Based on these assumptions, two types of real
adsorbed porosity (ads) can be calculated: intrinsic and scaled to bulk volume.

The intrinsic real adsorbed porosity is defined as the ratio of the adsorbed pore volume to the bulk volume of the organic
pore. It is mathematically expressed as,
! !
!!_!"# ! !!" !!!"# !_!" !!" !!!"#
!!"_! = = ! ! = ! (10)
!!_!" !!!" _!" !!"

The real adsorbed porosity scaled to the bulk volume of the composite system is obtained when the adsorbed pore volume is
related to the bulk volume of the shale rock. It can be expressed as follows,
!
!!_!"# !" !!" !!!"# !_!"
!"#_! = = (11)
!!_!! !!_!!

where !" is the mean of the pore size distribution of the shale, !"# and _!" are the thickness of the adsorbed layer and
length of organic pore corresponding to !" , respectively.

Adding Eq. (11) to Eq. (1) leads to the quintuple porosity of the shale as follows:

!! = ! + !"# + !"#_! + ! + !! (12)

In Eq. (12), org refers to the void space available within an organic pore to store free gas.

From Eq. (12), porosity ratios for organic, adsorbed, natural and hydraulic fractures are defined as follows,

!"# !"#_! ! !!
!"# = ; !"#_! = ; ! = ; !! = (13)
!! !! !! !!

where,

vorg = organic porosity ratio, fraction.


vads_c = real adsorbed porosity ratio, fraction.
8 SPE 165681

v2 = partitioning coefficient of natural fracture porosity ratio, fraction.


vhf = hydraulic fracture porosity ratio, fraction.

Material Balance Observations

Several researchers (Kucuk et al, 1978; Lewis, 2005; Mattar et al., 2008; Moghadam et al., 2009; Mengal and Wattenbarger,
2011; Cokar et al., 2012; Williams-Kovacs, 2012) have discussed the importance of the adsorption/desorption process on the
estimation of the Original Gas in Place (OGIP) in shale formations, and have developed mathematical models to include the
sorption mechanism in material balance calculations. These models try to capture the several storage mechanisms present in
shales and quantify their contribution to the total gas accumulation.

Kucuk et al. (1978) stated that gas is held in Devonian shales of the Appalachian basin in two ways (1): free gas in fractures,
and (2) adsorbed gas. They presented material balance plots for wells 6630 and 6654 in Lincoln County, West Virginia.
These plots are used to show that P/Z vs. cumulative gas production data for shale gas reservoirs deviate from the
conventional material balance for conventional gas reservoirs. They concluded that in Devonian shales, this deviation is an
indication of another storage mechanism: adsorption.

Figures 13 and 14 show the original data presented by Kucuk et al. (1978). The first straight line (black solid line) refers to
the early production of shales coming from free or compressed gas. Extrapolation of this straight line to P/Z equals to 0 leads
to an estimate of the initial free gas in place. Deviation from the first straight line corresponds to the initiation of the
desorption process which continues as indicated by the second straight line (green solid line). The interpretation developed by
Kucuk et al. (1978) only describes storage of free and adsorbed gas.

To the best of our knowledge there are not references in the literature indicating that a material balance plot can be used to
identifiy the contribution of diffusion from kerogen to the total cumulative gas production of shale gas reservoirs. Here we
use the same data presented by Kucuk et al. (1978) to identify this additional mechanism. The latter part of the curves in
Figures 13 and 14 shows deviation from the second to a third straight line. This behavior is an indication that in addition to
the adsorption/desorption process there is another source of gas in action.

The third straight line (orange solid line) in Figures 13 and 14 presents our interpretation of the data (this straight line was
not shown in Fucuk et al.s study). We associate the third straight line with a diffusion process within kerogen, which acts as
gas support to the total gas production. The dissolved gas within kerogen starts to diffuse from the solid kerogen as pressure
decreases and reaches a critical pressure for diffusion. The identification of the beginning of the third straight line in a P/Z vs.
cumulative production plot can be used as an approximation to estimate the critical pressure for gas diffusion. However, it is
highly recommendable to determine this parameter through laboratory experiments as diffusion is a slow process, and it
could take a long period of time to see its effects on a P/Z vs. cumulative production plot.

Mathematical Modeling of Shale Reservoirs

In this section, a mathematical description of shale gas reservoirs is introduced. The porosity systems are connected to each
other in the way established in Figure 10. For each porosity system considered in this model, its corresponding transport
equation is written. The model is derived based on the following assumptions and considerations,

The shale reservoir is made out of organic and inorganic matter, natural and hydraulic fractures
There is viscous flow in the fracture system
There is viscous flow, slip flow and Knudsen diffusion in organic and inorganic matter
There is adsorption/desorption from internal surfaces of organic pores
There is diffusion from solid organic matter (solid kerogen)
Single gas flow
There is stress-dependent permeability in the fracture network (natural and hydraulic fractures)
There is transient fluid transfer from matrix to natural fractures, and from natural to hydraulic fractures

Fluid transfer under transient conditions is used instead a pseudosteady state formulation as suggested by Brown et al. (2009).
These authors indicated that in general, transient models are more appropriate for reservoirs that exhibit low matrix
permeability values such as shale gas reservoirs. This statement resulted from the analysis of pressure and derivative
responses carried out in Brown et al.s work as shown in Figure 15. In our study we follow de Swanns (1976) transient
formulation. However, there are also other models available in the literature (Kazemi, 1969; Serra et al., 1983) which can be
considered for the analysis. Modifications to our proposed equations are required in order to include another transient model.
SPE 165681 9

The formulation presented in this paper includes stress dependency of fracture permeability. We use an equation proposed by
Jones (1975) to account for the effects of closure of fractures (natural and hydraulic fractures) when reservoir pressure
decreases and effective stress increases. However, other models could also be considered (Walsh, 1981; Gutierrez et al.,
2000; Raghavan and Chin, 2004; Chipperfield et al., 2007).

Hydraulic fractures

Viscous flow of free gas modeled by Darcys law is the assumed dominant transport mechanism in hydraulic fractures.
Hydraulic fractures deliver gas production to the well and they are fed by natural fractures according to the following
equation,

!!! !!! !! !! !
!! !! + ! = !! !!! !! (14)
! ! ! ! !"

where Qf refers to the mass transfer function between hydraulic and natural fractures. Based on the transient dual porosity
formulation developed by de Swaan (1976), we developed the mathematical expression for hydraulic fractures mass transfer
function as follows,

! ! !!!! !!
! !! , = !" ! ! (15)
!!! ! !!!

where quf refers to the flow caused by a unitary pressure loss at the surface, which is defined based on the theory of heat flow
in solids (Carslaw and Jaeger, 1959) and modified from de Swaans work,

!!
!" = ! !" (16)
!

At a new reservoir pressure Pi+1, the stress sensitive hydraulic fracture permeability is calculated using Jones equation,
!
!!!"!! ! !"# !!!!"!! !!"# !!
= (17)
!!!! !"# !!!!! !!"# !!

where Phf is the effective stress on the hydraulic fractures, and Ph is an apparent healing pressure that Jones (1975)
determined to be in the order of 40,000 psi from carbonate cores and from cores built with Portland cement. The validity of
the equation and the value of Ph have to be established for individual shale reservoirs.

Natural fractures

As in the case of hydraulic fractures, gas transport within natural fractures is modeled by considering viscous flow and
Darcys law. Based on our model (Figure 10), natural fractures are connected to both hydraulic fractures and production well
while both organic and inorganic matrices feed natural fractures. Consequently, our mass balance equation for natural
fractures is,

!! !! !! !! !! !! !
! ! ! + ! + !"#$ = ! !" ! (18)
! ! ! ! ! ! !"

where Qm is the mass transfer function between natural fractures and inorganic matrix, and Qorgf is the mass transfer function
between natural fractures and organic matrix. We adapt the transfer functions from original work by de Swaan (1976) as
follows,

! ! !!! !!
! ! , = !" ! ! (19)
!! ! !!!

and,

! ! !!! !!
!"#$ ! , = !"#$% ! ! (20)
!! ! !!!
10 SPE 165681

Inorganic (qum) and organic (quorgf) flow rates caused by unitary pressure loss at the surface are given by,

!!"
!" = ! !" (21)
!

and,

!!"#$
!"#$% = !"# !" (22)
!

where kam and kaorg are the apparent permeability of the inorganic and organic matrix, respectively. The apparent permeability
used in this paper follows the definition given by Civan (2010) and Swami and Settari (2012). These 2 definitions are
equivalent as shown in Appendix A.

For natural fractures, stress-dependent permeability is also defined by Joness (1975) expression,

!
!!!!! ! !"# !!!"!! !!"# !!
= (23)
!!" !"# !!!! !!"# !!

where Pf is the effective stress on natural fractures, and Ph is the apparent healing pressure determined by Jones (40,000 psi).
Depending on data availability different values of Ph can be used in equations 17 and 23.

Inorganic matrix

Inorganic matrix, also called as mineral matrix by Moghanloo et al. (2013) due to its composition of clay, silica and other
minerals, exhibits small size pores. As a consequence, viscous flow is not necessarily the dominant gas transport mechanism.
Additional mechanisms such as slip flow and Knudsen diffusion have to be considered. Some researchers (Swami and
Settari, 2012; Swami 2013) argue that inorganic pores are in the range of micropores, which lead to the fact that flow within
inorganic matrix can be represented by viscous flow. In our model we determine if viscous flow is or not valid based on the
Knudsen limits presented in Table 1. Consequently we present a general equation for the inorganic matrix that handles both
Darcy and non-Darcy flow through the use of an apparent permeability formulation (Civan, 2010; Swami and Settari, 2012).

Figure 10 shows that inorganic matrix is connected to natural fractures and is fed by the organic matrix. Therefore, the mass
balance equation for inorganic matrix is,

!!" !! !! !! !
! ! + !"#$ = ! !" ! (24)
! ! ! ! !"

where Qorgm is the mass transfer function between inorganic and organic matrices, defined as follows,

! ! !!! !!
!"#$ ! , = !"#$% ! ! (25)
!! ! !!!

The flow due to unitary pressure loss at the surface (quorgm) is,

!!"#$
!"#$% = !"# !" (26)
!

Pores in organic matrix and sorption

A well-developed pore network within organic matter (kerogen) has been identified by the use of SEM images (Ruppel and
Loucks, 2008; Wang and Reed, 2009; Ambrose et al., 2010; Loucks et al., 2012). The pores sizes are in the range of
nanopores. At nano-scale, the continuum and no slip flow break down. The gas transport deviates from viscous flow
behavior. Therefore, non-Darcy flow mechanisms are taken into account: slip flow and Knudsen diffusion. These additional
mechanisms are mathematically considered in the definition of apparent permeability (Javadpour, 2009; Civan, 2010; Swami
and Settari, 2012).

Based on the model we developed in Figure 10, organic matrix is connected to natural fractures and inorganic matrix. The
mass balance equation for organic matrix is,

!!"#$ !! !! !! ! !
!"# !"#$ !"#$ = !"# !"#$ !"# + !"# !"#_! ! (27)
! ! ! ! !" !"
SPE 165681 11

!
where !"#_! ! refers to the net rate of desorption that takes place on the surfaces of organic pores. Note that the mass
!"
exchange between free and adsorbed gas is limited by the real adsorbed porosity. This is because the adsorption/desorption
process only takes place in the internal surfaces of the organic pores (void space available for desorption), which is quantified
by the proposed real adsorbed porosity in this study. The real adsorbed porosity in equation 27 can change or remain constant
with time depending on the on-going desorption and adsorption process.

The surface area available in the organic matter for adsorption/desorption is significant and hence the amount of adsorbed gas
has to be taken into account in the analysis.

It is a common practice to use Langmuir isotherm to model shale gas reservoirs. This equation was developed under
equilibrium conditions, i.e. adsorption rate is equal to desorption rate. However, Do and Wang (1998) state that equilibrium
isotherms do not properly represent the interaction between free and adsorbed gas in complex and heterogeneous reservoirs
such as shales. Subsequently some researchers (Fathi and Akkutlu, 2009; Fathi, 2010; Akkutlu and Fathi, 2011; Shabro et al.,
2011; Shabro, 2012) have suggested that non-equilibrium adsorption should be implemented in modeling of shale reservoirs.

A discrepancy exists among researchers on the way that non-equilibrium conditions are considered. Fathi (2010) argues,
based on Do and Wangs observations, that desorption is much slower than adsorption. On the other hand, Shabro et al.
(2011) assumed that desorption surpasses adsorption during production. Based on that assumption, they developed a
formulation to model gas desorption at non-equilibrium conditions. In this paper, we define the net rate of desorption that
takes place on the surfaces of organic pores following Fathi and Akkutlus (2009) approach,

!!!
= !"# !"!# ! ! !"# ! (28)
!"

Diffusion from kerogen

Recently, dissolved gas in the solid kerogen has been considered as another source of gas in shale porous media. Some
researchers (Javadpour, 2007; Javadpour, 2009; Shabro et al., 2011; Swami and Settari, 2012; Shabro et al., 2012; Swami,
2013; Moghanloo et al., 2013) have conducted studies to quantify the amount of gas stored as dissolved gas in the solid part
of the organic matter. They have found that significant amount of gas is stored in this state in shale gas reservoirs. They have
concluded that dissolved gas should be considered for the calculation of gas in place in shale reservoirs. As gas desorption
takes place, gas concentration at kerogen surfaces decreases and causes a concentration gradient between the kerogen body
and kerogen surfaces. This concentration difference leads to gas diffusion from solid kerogen to kerogen surfaces.

To account for gas diffusion from solid kerogen, the 1D diffusivity equation in radial coordinates can be written as,

!!! ! ! !!!
= ! (29)
!" ! !" !"

Eqs. (28) and (29) are coupled by calculating gas concentration in the kerogen body and the concentration in the kerogen
surfaces.

The proposed mathematical formulations described above are being used to build a fully implicit numerical simulator from
scratch (Lopez and Aguilera, 2013). This will allow visualizing the effects of each of the mechanisms presented in this paper.

Production Variability

Shale petroleum reservoirs are very heterogeneous rocks with multiple porosities and storage mechanisms as discussed
above. All of these porosities have to be taken into account while calculating values of OPIP. Otherwise the results will be
very pessimistic. The same observation applies to production rates and recoveries. In spite of their complexity, there are
indication that cumulative gas distribution from shale gas reservoirs is more homogeneous than the distribution of tight gas
reservoirs as shown in the comparison presented in Figures 16 and 17.

The six curves in Figure 16 show production variability plots for different operating companies in the Barnett shale. If the
shales were completely homogeneous all the curves would fall in the dashed 45-degree straight line. The six curves
separation from the dashed straight line indicates a certain amount of heterogeneity (Aguilera, 2013).

Figure 17 is a cumulative production variability plot for the Nikanassin tight gas formation in six different areas of the Deep
Basin (Western Canada Sedimentary Basin) (Gonzalez and Aguilera, 2013). The data show that as the fracture density
decreases the curvature and thus the level of heterogeneity increases.
12 SPE 165681

The separation from the 45-degree dashed straight line is more significant in Figure 17 than in Figure 16 indicating a larger
degree of heterogeneity in tight gas reservoirs as compared with shale reservoirs. The result is surprising but corroborated by
actual production data. At this time there is not enough production data from tight and shale oil reservoirs to reach a
conclusion with respect to the distribution of their cumulative oil production.

Conclusions

1. A quintuple porosity system for shale gas reservoirs has been introduced. This system accounts for gas stored and
trapped as (1) adsorbed gas in kerogen in a real adsorbed porosity, (2) free gas in nonorganic matrix porosity, (3)
free gas in microfracture and slot porosity, (4) free gas in hydraulic fractures, and (5) free gas in the pore network
developed within kerogen. The methodology developed in the paper also accounts for gas dissolved in solid
kerogen.
2. A real adsorbed porosity has been developed based on the strict definition of porosity, i.e., ratio of pore volume to
bulk volume. Mathematical expressions for both intrinsic and scaled to bulk volume adsorbed porosities are
presented in this paper.
3. P/Z vs. cumulative production data from Devonian shale wells have been used to identify the adsorption process in
shale gas reservoirs. These data have also been used to identify an additional flow mechanism: diffusion from solid
kerogen.
4. A mathematical model for shale gas reservoirs modeling has been proposed. A transport equation is presented for
each porosity component of the composite system.
5. Fractional production variability plots show that the degree of production heterogeneity in shales is smaller than in
tight gas formations. This result is surprising due to the rock complexity associated with shale gas reservoirs.

Acknowledgements

The support of ConocoPhillips, the Natural Sciences and Engineering Research Council of Canada, Alberta Innovates Energy
and Environment Solutions, and the Schulich School of Engineering at the University of Calgary are gratefully
acknowledged. We thank the GFREE research team at the University of Calgary for their continued helps and support. We
also extend our gratitude to all the researchers who have been analyzing shale gas reservoirs to help develop a proper
understanding of the physics associated with these types of reservoirs.

Nomenclature

A= Area of the medium, m2


cg = Gas compressibility, Pa-1
ct = Total compressibility, Pa-1
Ca = Adsorbed gas concentration, kg/m3
Camax = Maximum monolayer adsorbed gas concentration, kg/m3
Ck = Gas concentration in kerogen bulk, kg/m3
D= Knudsen diffusion constant, m2/s
Dm = Matrixs Knudsen diffusion constant (accounting for porosity and tortuosity), m2/s
Dk = Diffusion constant for gas diffusion through solid kerogen, m2/s
F= Gas slippage factor, dimensionless
GsL = Langmuir storage capacity, scf/ton
Gads = Total gas adsorbed, kg
hads = Thickness of adsorbed layer, nm
hads_eq = Equivalent thickness of the adsorbed layer when only one pore is considered, nm
ka = Apparent permeability, m2
kads = Equilibrium constant for adsorption, kg/Pa-m2-s
kD = Darcy permeability, m2
kdes = Equilibrium constant for desorption, kg/m2-s
khf = Hydraulic fracture permeability, m2
km = Matrixs Darcy permeability (accounting for porosity and tortuosity), m2
L_op = Length of organic pore, m
P= Pressure, Pa
P = Effective stress, Pa
Ph = Apparent healing pressure determined by Jones (1975), Pa
Pk = Pressure of dissolved gas in solid kerogen, Pa
PL = Langmuir pressure, Pa
= Flow rate per unit rock volume, m3/s
SPE 165681 13

Q= Flow rate, m3/s


qL = Langmuir volume, m3/kg
T= Temperature, K
M= Molecular mass, kg/kmol
m= Real gas pseudopressure, Pa/s
n= Number of nanopores in the organic matter
NA = Avogadros number, 6.02x1023 molecules/mol
Rg = Universal gas constant, 8314 J/kmol/K
Rh = Hydraulic radius, m
r= Radial distance of radial coordinate system, m
rads = Radius of equivalent sphere of adsorbed gas, nm
rop = Organic pore radius, m
Sads = Adsorbent surface area, m
vads_c = Real adsorbed porosity ratio, fraction
vesp = Methane specific volume, cm3/g
vhf = Hydraulic fracture porosity ratio, fraction
vorg = Organic porosity ratio, fraction
v2 = Partitioning coefficient of natural fracture porosity ratio, fraction
V= Volume of a medium, m3
Vb = Bulk volume, m3
Vb_op = Bulk volume of organic pore, m3
Vb_sh = Bulk volume of shale rock, m3
Vp = Pore volume, m3
Vp_ads = Adsorbed pore volume, m3
Vstd = Mole volume of gas at standard temperature (273.15 K) and pressure (101 325 Pa), 22.413x10 -3 m3/mol
Z= Gas compressibility factor

Greek Symbols

= Fraction of gas molecules coming back from the pore wall, dimensionless
= Difference operator
= Gas viscosity, Pa.s
b = Bulk density of shale, kg/m3
= Gas density, kg/m3
s = Sorbed phase density, kg/m3
sc = Gas density at standard temperature (273.15 K) and pressure (101 325 Pa), kg/m3
= Porosity, fraction
ads = Real adsorbed porosity, fraction
ads_c = Real adsorbed porosity scaled to the bulk volume of the composite system, fraction
ads_i = Intrinsic real adsorbed porosity, fraction
= Tortuosity, dimensionless
! = De Swaans integration parameter

Subscripts

2= Natural fractures
hf = Hydraulic fractures
m= Matrix
org = Organic matter (kerogen)
sh = Shale (composite system)
u= Solution with unitary-step boundary condition

Acronyms

ERCB = Energy Resources Conservation Board (Alberta, Canada)


GFREE = Integrated geoscience (G), formation evaluation (F), reservoir drilling, completion and stimulation (R), reservoir
engineering (E), economics and externalities (EE)
OGIP = Original gas in place
OPIP = Original petroleum in place
14 SPE 165681

References

Aguilera, R., 1995, Naturally Fractured Reservoirs, PennWell Books, Tulsa, Oklahoma.

Aguilera, R., 2006, Effect of Fracture Compressibility on Oil Recovery From Stress-Sensitive Naturally Fractured Reservoirs, Journal of
Canadian Petroleum Technology, Volume 45, No. 12, December 2006.

Aguilera, R., 2010, A Method for Estimating Hydrocarbon Cumulative Production Distribution of Individual Wells in Naturally Fractured
Carbonates, Sandstones, Shale Gas, Coalbed Methane and Tight Gas Formations, Journal of Canadian Petroleum Technology, Volume 49,
No. 8, August 2010.

Aguilera, R., 2010, Flow Units: From Conventional to Tight Gas to Shale Gas Reservoirs, SPE Paper 132845 presented at the Trinidad and
Tobago Energy Resources Conference held in Port of Spain, Trinidad, 2730 June 2010.

Aguilera, R., 2013, Flow Units: From Conventional to Tight Gas to Shale Gas to Tight Oil to Shale Oil Reservoirs, SPE Paper 165360
presented at the SPE Western Regional & AAPG Pacific Section Meeting, 2013 Joint Technical Conference held in Monterey, CA, April
1925, 2013.

Akkutlu, I. Y., and Fathi, E., 2011, Multi-scale Gas Transport in Shales with Local Kerogen Heterogeneities, SPE paper 146422 presented
at the SPE Annual Technical Conference and held in Denver, CO, 30 October2 November, 2011.

Ambrose, R. J., Hartman, R. C., Diaz-Campos, M., Akkutlu, I. Y., and Sondergeld, C. H., 2010, New Pore-Scale Considerations for Shale
Gas in Place Calculations, SPE paper 131772 presented at the SPE Unconventional Gas Conference held in Pittsburgh, PA, February 23-
25, 2010.

Andrade, J., Civan, F. and Devegowda, D., Sigal, R. F., 2011, Design and Examination of Requirements for a Rigorous Shale Gas
Reservoir Simulator Compared to Current Shale Gas Simulators, SPE paper 144401 presented at the SPE North American Unconventional
Gas Conference held in the Woodlands, Texas, 14-16 June 2011.

Beliveau, D., 1993, Honey, I Shrunk the Pores!, Journal of Canadian Petroleum Technology, Volume 32, No. 8, October 1993.

Brown, G. P., Dinardo, A., Cheng, G. K., and Sherwood, T. K., 1946, The Flow of Gases in Pipes at Low Pressures, Journal of Applied
Physics, October, pp. 802-813.

Brown, M., Ozkan, E., Raghavan, R., and Kazemi, H., 2011, Practical Solutions for Pressure-Transient Responses of Fractured Horizontal
Wells in Unconventional Shale Reservoirs, SPE Reservoir Evaluation and Engineering, December, pp. 663-676.

Carslaw, H. S., and Jaeger, J. C., 1959, Conduction of Heat in Solids, Oxford U. Press, New York.

Cheadle, B. A. and Jiang, P., 2013, Born this Way Inherited Heterogeneity and Microporosity Modalities in Hybrid Reservoirs of the
Upper Cretaceous Colorado Group, Western Canada Foreland Basin, presented at the CSEG-CSPG-CWLS GeoConvention on Geoscience
and Engineering partnership.

Chornaya, V., Lipatov, Y., Todosijchuk, T., and Menzheres, G., 2002, Effect of Temperature on the Structure of Adsorption Layers
Formed by Adsorption of Polymers in the Transition Region from Dilute to Semidilute Solutions, Journal of Colloid and Interface Science,
July, pp. 36-43.

Civan, F., 2010, A Review of Approaches for Describing Gas Transfer through Extremely Tight Porous Media, American Institute of
Physics, June, pp. 53-58.

Civan, F., 2010, Effective Correlation of Apparent Gas Permeability in Tight Porous Media, Transport in Porous Media, Volume 82, pp.
375-384.

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, November, p.208-223.

Darabi H., Ettehad, A., Javadpour, F., and Sepehrnoori K., 2012, Gas flow in ultra-tight shale strata, Journal of Fluid Mechanics, Volume
710, pp. 641-658.

deSwaan O., A., 1976, Analytic Solutions for Determining Naturally Fractured Reservoir Properties by Well Testing, SPE Journal, June,
pp. 117-122.
SPE 165681 15

Ertekin, T., King, G. R. and Schwerer, F. C., 1986, Dynamic Gas Slippage: A Unique Dual-Mechanism Approach to the Flow of Gas in
Tight Formations, SPE Formation Evaluation, February, pp. 4352.

Fathi, E., and Akkutlu, I. Y., 2009, Matrix Heterogeneity Effects on Gas Transport and Adsorption in Coalbed and Shale Gas Reservoirs,
Transport in Porous Media, Volume 80, pp. 281-304.

Fathi, E., 2010, Matrix Heterogeneity Effects on Fluid Transport in Porous Medium Using Perturbation Theory, PhD thesis, Mewbourne
School of Petroleum and Geological Engineering, University of Oklahoma.

Gonzalez, L. and Aguilera, R., 2013, Effect of Natural Fracture Density on Production Variability of Individual Wells in the Tight Gas
Nikanassin Formation, paper CSUG/SPE 149222-PP presented at the Canadian Unconventional Resources Conference held in Calgary,
Alberta, Canada, 1517 November 2011. Journal of Canadian Petroleum Technology, March, 2013, pp. 131-143.

Hudson, J. D., 2011, Quad-Porosity Model for Description of Gas Transport in Shale-Gas Reservoirs, MSc thesis, Mewbourne School of
Petroleum and Geological Engineering, University of Oklahoma.

Hudson, J. D., Civan, F., Michel-Villazon, G., Devegowda, D., and Sigal, R., 2012, Modeling Multiple-Porosity Transport in Gas-Bearing
Shale Formations, SPE paper 153535 presented at the SPE Latin American and Caribbean Petroleum Engineering Conference held in
Mexico City, Mexico, April 16-18, 2012.

Javadpour, F., Fisher, D., Unsworth, M., 2007, Nanoscale Gas Flow in Shale Gas Sediments, Journal of Canadian Petroleum Technology,
Volume 46, No. 10, October 2007.

Javadpour, F., 2009, Nanopores and Apparent Permeability of Gas Flow in Mudrocks (Shales and Siltstone), Journal of Canadian
Petroleum Technology, Volume 48, No. 8, August 2009.

Jones, Jr., F. O., 1975, A Laboratory Study of the Effects of Confining Pressure on Fracture Flow and Storage Capacity in Carbonate
Rocks, Journal of Petroleum Technology, January, pp. 21-27.

Kucuk, F., Alam, J., and Streib, D. L., 1978, Reservoir Engineering Aspects and Resource Assesment Methodology of Eastern Devonian
Gas Shales, Science Applications, Inc., October.

Lopez, B. and Aguilera, R., 2013, Construction of a Numerical Simulation Model for Handling Quintuple Storage in Shale Gas Reservoirs,
in preparation for presentation at SPE meeting.

Loucks, R. G., Reed, R. M., Ruppel, S. C., and Hammes, U., 2010, Preliminary classification of matrix pores in mudrocks: Gulf Coast
Association of Geological Societies Transactions, v. 60, pp. 435-441.

Loucks, R. G., Reed, R. M., Ruppel, S. C., and Hammes, U., 2012, Spectrum of pore types and networks in mudrocks and a descriptive
classification for matrix-related mudrock pores, AAPG Bulletin, v. 96, no. 6, pp. 10711098.

Michel, G. G., Sigal, R. F., Civan, F., Devegowda, D., 2011, Parametric Investigation of Shale Gas Production Considering Nano-Scale
Pore Size Distribution, Formation Factor, and Non-Darcy Flow Mechanisms, SPE paper 147438 presented at the SPE Annual Technical
Conference and Exhibition held in Denver, CO, 30 October2 November, 2011.

Moghanloo, R. G., Javadpour, F., and Davudov, D., 2013, Contribution of Methane Molecular Diffusion in Kerogen to Gas-in-Place and
Production, SPE paper 165376 presented at the SPE Western Regional & AAPG Pacific Section Meeting, 2013 Joint Technical Conference
held in Monterey, CA, April 1925, 2013.

Ozkan, E., Raghavan, R., Apaydin, O. G., 2010, Modeling of Fluid Transfer from Shale Matrix to Fracture Network, SPE paper 134830
presented at the SPE Annual Technical Conference and Exhibition held in Florence, Italy, September 19-22, 2010.

Wischnewski, B., peace software, Berlin, 2013, Calculation of thermodynamic state variables of methane, Methane at normconditions,
http://www.peacesoftware.de/einigewerte/methan_e.html

Sakhaee-Pour, A., and Bryant, S. L., 2012, Gas Permeability of Shale, SPE Reservoir Evaluation and Engineering, August, pp. 401-409.

Schettler, P. D., Parmely, C. R., and Lee, W. J., 1989, Gas Storage and Transport in Devonian Shales, SPE Formation Evaluation,
September, pp. 371376.

Shabro, V., Torres-Verdn, C., and Javadpour, F., 2011, Numerical Simulation of Shale-Gas Production: From Pore-Scale Modeling of
Slip-Flow, Knudsen Diffusion, and Langmuir Desorption to Reservoir Modeling of Compressible Fluid, SPE paper 144355 presented at the
SPE North American Unconventional Gas Conference and Exhibition held in The Woodlands, TX, June 14-16, 2011.
16 SPE 165681

Shabro, V., Torres-Verdn, C., and Sepehrnoori, K., 2012, Forecasting Gas Production in Organic Shale with the Combined Numerical
Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from Kerogen Surfaces, and Advection in Nanopores, SPE paper 159250
presented at the SPE Annual Technical Conference and Exhibition held in San Antonio, TX, October 8-10, 2012.

Sigal, R. F., Akkutlu, I. Y., Kang, S. M., Diaz-Campos, M., and Ambrose, R., 2013, The Laboratory Measurement of the Gas Storage
Capacity of Organic Shales, shale.ou.edu/Content/Member%20Area/Papers/GasStorage.pdf (last entered May 14, 2013).

Swami V., and Settari, A. T., 2012, A Pore Scale Gas Flow Model for Shale Gas Reservoir, SPE paper 155756 presented at the Americas
Unconventional Resources Conference held in Pittsburgh, PA, June 5-7, 2012.

Swami V., 2012, Shale Gas Reservoir Modeling: From Nanopores to Laboratory, SPE paper 1630365-STU prepared for the presentation at
the SPE International Student Paper Contest at the SPE Annual Technical Conference and Exhibition held in San Antonio, TX, October 8-
10, 2012.

Swami, V., 2013, Development of a Quad Porosity Numerical Flow Model for Shale Gas Reservoirs, MSc thesis, Schulich School of
Engineering, University of Calgary.

Tran, T., Sinurat, P., and Wattenbarger, R. A., 2011, Production Characteristics of the Bakken Shale Oil, SPE paper 145684 presented at
the SPE Annual Technical Conference and Exhibition held in Denver, CO, 30 October2 November, 2011.

Vasina, E. N., Paszek, E., Nicolau, D. V. Jr., and Nicolau, D. V., 2009, The BAD project: data mining, database and prediction of protein
adsorption on surfaces, The Royal Society of Chemistry, April, pp. 891-900.

Wang, F. P. and Reed, R. M., 2009, Pore networks and Fluid Flow in Gas Shales, SPE paper 124253 presented at the SPE Annual
Technical Conference and Exhibition held in New Orleans, Louisiana, October 4-7.

Yan, B., Wang, Y., and Killough, J. E., 2013, Beyond Dual-Porosity Modeling for the Simulation of Complex Flow Mechanisms in Shale
Reservoirs, SPE paper 163651 presented at the SPE Reservoir Simulation Symposium held in The Woodlands, TX, February 18-20, 2013.

Appendix A

Civan (2010) derived the following apparent permeability expression for a real gas,

! !!! ! !!!
! !! !!! ! !"
! = + 1 (A-1)
!!! !!! !" !!! !" ! !"

Darcys permeability (kD) for a circular tube is expressed using the Hagen-Poiseuilles equation,
!
!!
! = (A-2)
!

It is known that Knudsen diffusitivity (D) is given by (Roy et al., 2003),

!!! !!! !
= (A-3)
! !"

Inserting (A-2) and (A-3) into (A-1) leads to,

! ! ! ! !"
! = ! + 1 (A-4)
!! !! !!! !" ! !"

Gas density () is given by the real gas equation of state,

!"
= (A-5)
!!! !

Substituting (A-5) into (A-4) results in,

! ! ! ! !"
! = ! + 1 (A-6)
!! !! ! ! !"
SPE 165681 17

The above equation can be rearranged as,

! ! ! ! !"
! = ! + (A-7)
!! !! ! ! !"

Gas compressibility is defined as,

! ! !"
! = (A-8)
! ! !"

Substituting (A-8) into (A-7) leads to,

! !
! = ! + ! (A-9)
!! !!

Darcy permeability and Knudsen diffusivity for the porous medium are expressed as,

!
! = ! (A-10)
!

!
! = (A-11)
!

Substitution of (A-10) and (A-11) into (A-9) gives,

! = ! ! + ! (A-12)

Thus the apparent permeability given by Eq. (6) in the main body of the manuscript (Swami and Settari, 2012) is the same as
Eq. (A-12) developed from Civans (2010) Eq. (A-1).

Appendix B

Vasina et al. (2009) presented a method for predicting the thickness of adsorbed protein films. Here we adapt their
methodology to calculate the thickness of a methane adsorbed layer (hads) required for computation of the real adsorbed
porosity developed in this paper.

The methane molecule is approximated by a sphere and its volume is calculated based on the methane specific volume. For a
real gas, the specific volume is given by,

! !"#
!"# = = (B-1)
! !"

At 21 C and 28 MPa, the gas compressibility factor Z is equal to 0.913 according to the peace software calculation. For
these conditions, the specific volume of methane is,
!
!.!"# !.!"# !" !"# !"#$ !"#.!" ! !"!
!"# = ! = 4.97 (B-2)
!".!"# !"# !" !"# !

The methane volume is defined as,


!!"# !
= (B-3)
!!

where NA is the Avogadros number. Upon substitution of the required values into (B-3), the methane volume is,

!"! !
!.!" !
!".!"# !"#
= !

= 1.32410!!! ! (B-4)
!.!"!!"!"
!"#

If the methane volume is expressed in cubic nanometers, it is equal to,

!! !! !! !"!
= 1.32410!!! ! = 0.1324 ! (B-5)
!""! !"! !!!"!! ! !!
18 SPE 165681

Since the methane molecules are approximated by a sphere, the radius of the equivalent sphere is,

! !
!! ! ! !.!"#$ !"! !
!"# = = = 0.316 (B-6)
!! !!

The thickness of a monolayer is the diameter of the equivalent sphere. Therefore, the thickness of the adsorbed layer is,

!"# = 2!"# = 2 0.316 = 0.632 (B-7)

Ambrose et al. (2010) suggested a typical value of 0.7 nm for the same pressure and temperature. However, they did not
show how their value was calculated.

Lipatov et al. (2002) developed a different approach, in this case for polymers, to calculate the thickness of a polymer
adsorption layer. We have adapted their method to calculate the thickness of an adsorbed layer of methane on organic matter
as follows,

!!"#
!"# = (B-8)
!!!"#

where Gads is the total gas adsorbed, is the density of the adsorbate and Sads is the adsorbent surface. If Langmuir adsorption
is assumed, the total gas adsorbed is,

!
!"# = !" ! ! ! = !" ! ! !" (B-9)
!!!!

Considering the values of Table 2, GL is calculated as,

!"# !"#$.!" !"#$ !"# !"#


! = 50 = 38.966 = 0.001216 (B-10)
!"# !"#$.!"!!!"# !"#$ !"# !"

The total gas adsorbed (Gads) is,

!" !" !"#


!"# = 0.717 2500 1 ! 1 0.001216 = 2.1802 (B-11)
!"# !! !"

For 1m3 of rock and total porosity of the shale equal to 0.06, the pore volume is,

! = !! !_!! = 0.06 1 ! = 0.06 ! (B-12)

Assuming a cylindrical geometry for organic pores to estimate the radius of each pore,

!!
= (B-13)
!!_!" !

where n is the number of pores. If it is assumed that there are 1x1015 nanopores in the rock, the radius of the pore is,

!.!" !!
= = 4.3710!! = 4.37 (B-14)
! !! !!!"!"

Based on the assumption of a cylindrical geometry for organic pores, the adsorbent surface area is,

!"# = 2!" _!" (B-15)

Considering the values of Table 2 leads to,

!"# = 2 110!" 4.3710!! 1 = 2.7510! ! (B-16)

The gas density () is given by the real gas equation of state,


!"
= (B-17)
!"#
SPE 165681 19

At 21 C and 28 MPa, the gas density is,


!
!".!"# !"# !" !"# ! !"
= ! !"# = 0.2012 = 201.2 (B-18)
!.!"# !.!"# !" !"! !!
!"#$ !"#.!" !

Finally, the thickness of methane adsorbed layer is,

!.!"#$ !"
!"# = !" = 3.94610!!" = 0.394 (B-19)
!"#.! !.!"!!"! !!
!!

In conclusion, according to the first method, the thickness of the methane adsorbed layer is approximately equal to 0.63 nm
whereas the second method calculates a thickness of the methane adsorbed layer equals to 0.39 nm. The difference between
these two values stems from the assumptions made for the development of each equation.

Table 1. Flow regime classification based on Knudsen number limits (after Schaaf and Chambre, 1961; and Civan, 2010)

Flow Regime Knudsen Number

Continuum flow <=0.001

Slip flow 0.001 0.1

Transition flow 0.1 10

Free-molecular flow >= 10

Table 2. Parameters used for the calculation of the thickness of adsorbed layer.

Parameter Value Units Source


Reservoir pressure, P 4061.06 psia Ambrose et al. (2010)
Langmuir volume, GSL 50 scf/ton Ambrose et al. (2010)
Langmuir pressure, PL 1150 psia Ambrose et al. (2010)
3
Methane densitySC, sc 0.717 kg/m peace software (2013)
3
Bulk density, b 2500 kg/m Ambrose et al. (2010)
Net pay, h 1 m Beliveau (1993)
2
Area, A 1 m Beliveau (1993)
Total porosity of shale, ads 0.06 m Ambrose et al. (2010)
Nano pore length, L_op 1 m Beliveau (1993)
20 SPE 165681

Figure 1. Pore type classes. (A) Spectrum of pore types occurring within shale gas systems. General pore types include interparticle
and intraparticle pores associated with mineral matrix, organic-matter pores associated with organic matter, and fracture pores that
crosscut matrix and grains (Loucks et al., 2012). The figure shows examples of common pore types but they are not meant to
indicate formal subtypes or terminology. Loucks et al. present fracture pores in this figure only to emphasize their importance.
However, they do not considered them to be matrix-related pores. (B) The shale gas system pore classification ternary diagram was
modified by Loucks et al. (2011) from Loucks et al. (2010). If a pore network is dominated by 50% or more of one pore type, it
assumes the name of that pore type. If no one pore type dominates, then it is a mixed-pore network.
SPE 165681 21

Figure 2. Well-defined pore structure within organic matter (kerogen). Upper part: Colorado shale in the Western Canada
Sedimentary Basin (courtesy of C. D. Rokosh of the ERCB, 2010). Lower part: Barnett shale (provided by R. Reed to Ruppel and
Loucks, 2008).
22 SPE 165681

Figure 3. Six exchange pathways between the different porosity components in shale reservoirs as defined by Civan (2012).

Organic matrix Inorganic matrix Fractures

Figure 4. Gas transport in series in organic-rich shale reservoirs presented by Akkutlu and Fathi (2011).

Figure 5. Effect of desorption on production velocity (Source: Shabro et al., 2011).


SPE 165681 23

Figure 6. Modeling of organic and inorganic porosity in parallel without interaction between them (Hudson, 2011). Upper part: Tank
representation of connectivity of each porosity component of the quadruple system in shale gas reservoirs. Lower part: 2D
visualization of the simulation model used by Hudson (2011). OP: organic porosity, IP: inorganic porosity, NF: natural fractures, HF:
hydraulic fractures, Prod: production well.

Figure 7. Contribution of each flow mechanism considered in the reservoir model developed by Shabro et al. (2012). Conventional:
only Darcy flow; Add slip flow: Darcy flow, slip flow and Knudsen diffusion; Add desorption: Darcy flow, slip flow, Knudsen
diffusion and Langmuir desorption; Add diffusion: Darcy flow, slip flow, Knudsen diffusion, Langmuir desorption and diffusion from
kerogen.
24 SPE 165681

Figure 8. Contribution of each component of the 'quad porosity' system defined by Swami (2013) as compared against experimental
data (Javadpour et al., 2007). An excellent agreement between simulated and experimental data was achieved.

(1) Adsorption

(5) Organic (2) Inorganic


Matter Matter
Portfolio QUINTUPLE
POROSITY &
DIFFUSION

(4) Natural (3) Hydraulic


Fractures Fractures

Figure 9. Quintuple porosity system including adsorption, organic and inorganic matter, natural and hydraulic fractures. Gas
diffusion from solid kerogen is considered as an additional storage mechanism.
SPE 165681 25

Figure 10. General gas transport formulation of the quintuple porosity system proposed in this study for shale gas reservoirs.

Figure 11. Simplifications of the general gas transport in shale reservoirs proposed in this paper. Red solid lines show the paths
followed for each of the simplifications. Upper half: Series gas transport. Lower half: Parallel gas transport defined by Hudson
(2011).
26 SPE 165681

Figure 12. Schematic of an organic pore showing the thickness of the adsorbed layer (hads) and its relation to the radius of the pore
(rop).

Figure 13. P/Z vs. cumulative production crossplot for Well No. 6630, Lincoln County, WV (modified from Kucuk et al., 1978). The
orange solid line presents our interpretation for an additional mechanism (dissolved gas in kerogen) added to the two storage
mechanisms (free and adsorbed gas) pointed out by Kucuk et al. (1978).
SPE 165681 27

Figure 14. P/Z vs. cumulative production crossplot for Well No. 6654, Lincoln County, WV (modified from Kucuk et al.,1978). The
orange solid line presents our interpretation for an additional mechanism (dissolved gas in kerogen) added to the previous two
storage mechanisms (free and adsorbed gas) pointed out by Kucuk et al. (1978).

Figure 15. Comparison of pseudosteady and transient fluid transfer from matrix to fracture (modified from Brown et al., 2009).
Impact of choice of dual-porosity model on pressure and derivative responses.
28 SPE 165681

100
BASE
Devon
Burlington
80
Encana
XTO
% CUMULATIVE GAS

Chief
60

40

20

0
0 10 20 30 40 50 60 70 80 90 100
% WELLS

Figure 16. Fractional production variability plot for the Barnett fractured shale in Texas for various operating companies (Source:
Aguilera, 2013).
































Figure 17. Fractional production variability plot for six Nikanassin tight gas areas of the Deep Basin of Canada (Source: Gonzalez
and Aguilera, 2013).

S-ar putea să vă placă și