Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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)

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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


Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022
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 3.8 7.6 11.4 15.2 19.
0
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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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)

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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
V 10.73159 519.67˚R
DOE/NETL, Alabama University, Tuscaloosa, Alabama. = ˚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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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

Downloaded from http://onepetro.org/SJ/article-pdf/17/01/219/2107568/spe-131772-pa.pdf by Algerian Petroleum Inst. user on 31 October 2022


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 ⎥
= 155.7 scff/ton. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-12)
= ⎢ −6 ⎥
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

You might also like