Вы находитесь на странице: 1из 67

UNIVERSITY OF OKLAHOMA

GRADUATE COLLEGE




UNCERTAINTIES IN SHALE GAS IN-PLACE CALCULATIONS: MOLECULAR SIMULATION APPROACH


A THESIS
SUMMITTED TO THE GRADUATE FACULTY
in partial fulfillment of the requirement for the
Degree of
MASTER OF SCIENCE



By
MERY DIAZ CAMPOS
Norman, Oklahoma
2010











UNCERTAINTIES IN SHALE GAS IN-PLACE CALCULATIONS: MOLECULAR SIMULATION APPROACH




A THESIS APPROVED FOR THE
MEWBOURNE SCHOOL OF PETROLEUM AND GEOLOGICAL ENGINEERING

BY


___________________________________
I. Yucel Akkutlu, Chair


___________________________________
Carl Sondergeld


___________________________________
Chandra Rai































Copyright by MERY DIAZ CAMPOS 2010
All Rights Reserved.
iv


ACKNOWLEDGEMENTS

First of all, I thank Dr. Yucel Akkutlu for supervising my research during these two years.
Without his new ideas it would not have been possible to develop the investigation to its
present stage. Thanks for the constant encouragement and the academic freedom he gave to
me during my research. Thanks for all his advice, both technical and personal. It was very
encouraging for me every time he showed such enthusiasm looking at the results of the
simulations as myself.
I would like to thank my committee members Drs. Chandra Rai and Carl Sondergeld for
their encouragements to start molecular-level investigations related to reservoir phenomena at
the University of Oklahoma.
The technical support I received during the development of this work has been
indispensable to improve the quality of the molecular dynamics simulations developed in this
work. In particular I would like to mention Dr. Alberto Striolo and Dr. Richard Sigal they deserve
much credit in generating an appropriate ensemble for the study developed in the present
thesis. Thanks to all the faculty of PE for sharing with me the enthusiasm of learning about this
new molecular level approach in order to understand some physical chemistry properties of
hydrocarbon fluids when confined in micro- and nano-porous media.
It was a pleasure for me to discover that Dr. Akkutlu and I were not the only ones
interested in the results of the molecular simulations involving nanopores, but also my family
and friends. I would like to thank my uncle and my mother, in particular, who always follow up
v

my performance with greatest interest. Thanks to my entire family, although being far away,
they have been my inspiration to continue learning. Thanks to my good friends Chau Diep,
Yoana Walschap, Pinar Akkutlu, Serita and all my Sooner teammates, for making me feel like
being at home.
The simulation studies were performed using parallel computing at the OU
Supercomputing Center for Education & Research (OSCER). I thank the director Dr. Henry
Neeman for his support of my studies.
My education and research has been supported by Devon Energy. I am grateful for this
support.
iv


TABLE OF CONTENTS

Page
Acknowledgements iv
Table of contents . iv
List of Tables vi
List of Figures vii
ABSTRACT ix
CHAPTERS
1. INTRODUCTION
1.1 Overview .. 1
1.2 Purpose of study .......... 6
2. MOLECULAR MODELING AND SIMULATION
2.1 Molecular Dynamics Simulation .. 9
2.2 Monte Carlo Simulation 11
2.3 Numerical Procedures for Gas Adsorption 13
2.4 Numerical Procedure for Methane Solubility in Nanopores 19
3. NUMERICAL RESULTS AND DISCUSSION
v

3.1 Methane Adsorption in Organic Pores using Molecular
Simulations.. 22
3.2 Methane Solubility Enhancement in Water confined to nano-scale pores using Molecular
Dynamics Simulations 32
4. CONCLUSIONS AND RECOMMENDATIONS
4.1 Conclusions ... 36
4.2 Recommendations ... 37
REFERENCES 38
NOMNECLATURE . 42
vi


LIST OF TABLES

Page
CHAPTER 1
Table 1.1- Typical TOC contents of North American Shale Gas Plays . 1
Table 2.1- Lennard-Jones Potential Parameters for Methane-Graphite System ... 17
Table 2.2- Lennard-Jones potential parameters for Methane-Water-Graphite System 22
Table 3.1- Methane density at different pressures. Comparison with the bulk density
values obtained using SUPERTRAPP under the same pressure and
temperature conditions. .. 28
Table 3.2- Calculated values of the Solubility Parameters l, L and Henrys Law Constant
k
H
at 75.5
o
C .. 37
Table 3.3- Effect on the solubility parameters l, L at 75.5
o
C by changing the wall-water
interaction. The pore width = 1.62nm . 40
Table 3.4- Shale properties, low and high sorption capacity, taken from (Ambrose et al.,
2010), including three different density phases for comparison . 41
Table 3.5- Free and total storage capacity for Shale A and B, at different adsorbed
density phase values . 43
vii


LIST OF FIGURES

Page
CHAPTER 1
Figure 2.1 Molecular simulation cell consisting of graphite walls and OPLS-UA methane
.. 14
Figure 2.2 Methane insertion at a vacant spot . 25
Figure 2.3 Total energy (left) and temperature (right) fluctuation over the time for a pure water
system in thermodynamic equilibrium in a nanopore at 75.5
o
C . 25
Figure 3.1- Number density (left) and discrete density (right) profiles for methane in 3.9 nm
graphite slit-pore at 176
o
F . 29
Figure 3.2 Number density (top) and discrete density (bottom) profiles for methane in 2.31 nm
graphite slit-pore at 176
o
F .. 27
Figure 3.3 Number density (left) and discrete density (right) profiles for methane in 1.14 nm
graphite slit-pore at 176
o
F .. 32
Figure 3.4 Number density (top) and discrete density (bottom) profiles for methane in 3.93nm
slit-pore as a function of temperature ... 33
Figure 3.5 Density profiles for methane in three different-size slit-pore 1.14, 2.3 and 3.8nm
using MC simulation ... 34
viii

Figure 3.6 Number density (top) and discrete density (bottom) profile for carbon dioxide in
3.8nm slit-pore as function of distance from the wall to the center of the pore, at p =
1980 psi .. 35
Figure 3.7 Estimated Ostwald coefficient, L (x10
3
), of methane in water confined to slit-pores
with varying sizes compared with the solubility in bulk water (top). Corresponding
SCF Methane/BBl water (down) at 200psi and 168F 38
Figure 3.8 Estimated Ostwald coefficient, L (x10
3
), of methane in water confined to 1.62 nm
pore with varying pore-wall wettability 40



ix


ABSTRACT
Shale gas reservoirs are a major source for the global energy market in the US. In 2008, dry
natural gas proved reserves attributable to shale reservoirs grew to a total of 32.8 tcf from 21.7
tcf in 2007 (Annual Energy Outlook 2010). On the other hand regarding natural gas production,
shale gas is currently and will be the largest contributor to its growth. EIA predicts growth from
2.7tcf to 6tcf during 2010 and 2035 (Long, G. 2010).
Shale gas-in-place calculation consists of a volumetric component representing
hydrocarbons stored in the pore space of the reservoir system as free gas. In addition, a surface
component is necessary to account for the gas physically adsorbed on large internal surface area
of the organic microporous matrix. A second volumetric component, which is not considered
within the calculation is gas dissolved in the formation water or liquid hydrocarbons. For such
reservoirs, free gas is typically quantified using volumetric method. Adsorbed gas has generally
been estimated from laboratory studies using equilibrium adsorption isotherms, under
laboratory conditions and requires a correction for the adsorbed gas fraction present under
reservoir temperature and pressure conditions. The amount of dissolved gas in formation water
has not been considered important, based on bulk solubility calculations.
The thesis is based on a series of theoretical studies using molecular modeling and
simulations to investigate physical chemical properties of fluids confined to organic micropores.
The amounts of adsorbed, dissolved and free natural gas are predicted using fluids in carbon slit
pores with varying size and pore-wall wettability under typical reservoir pressure and
temperature conditions. It is found that gas adsorption and solubility in liquid are enhanced due
to dominant pore wall effects. These effects are controlled to a varying degree by the micropore
x

size, temperature, pressure and fluid composition. The implication of these numerical
observations is that the sorbed gas in the formation may represent a significant portion of the
total gas reserves in unconventional reservoirs.
1


CHAPTER 1
INTRODUCTION

1.1 Overview
Production of natural gas from organic-rich shale makes up an increasingly greater portion of
the total production in North America. 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 to over 2,000 bscf, and has continued to grow (EIA,
2010). The main portion of the shale gas production is maintained by the large US shale plays
like the Barnett, while growth is due to new production coming on line in the Woodford,
Haynesville, Marcellus, Eagle Ford, and Horn River, Avalon to name a few.
Natural gas in these reservoirs is stored by two mechanisms, free gas and adsorbed gas.
The adsorbed gas is stored mainly on the internal surfaces of the organic material. Table 1.1
shows average total organic content (TOC) by weight percent of the shales (Pollastro et al, 2007;
Jenden et al., 1993; Mancini et al., 2006; Comer and Hinch, 1987; Reynolds and Munn, 2010).

Shale or Play

TOC (wt %)
Barnett 4%
Marcellus 1-10%
Haynesville 0-8%
Horn River 3%
Woodford 5%

Table 1.1- Typical TOC contents of North American Shale Gas Plays
2

When conducting a reservoir study on a natural gas field, one of the primary concerns is
the estimation of initial gas-in-place. The estimate is the basis for disclosure of gas reserves and
it is important for reservoir engineering analysis such as gas production forecast. Tank-type and
multi-dimensional (simulation-based) material balance calculations are common industry
approaches in predicting the gas-in-place when sufficient field performance data are available.
For disclosure purposes, a deterministic volumetric method, in which a single average
value is selected for each parameter in the reserves calculations, is most commonly used in
North America. Probabilistic methods, on the other hand, are increasingly used worldwide and
give the ability to describe the full range of values for each parameter in order to somewhat
reflect spatial variability in the parameters and structural intricacies in reservoir architecture. In
the present research it is shown that any volumetric approach for shale gas reserve estimation
has added complexity, because in shale the natural gas (mostly methane) exists in different
thermodynamic states namely: adsorbed, absorbed (or dissolved) and free gas, and that an
accurate estimation of the gas pore volume in shale reservoirs should not be considered
independent of these states.
A simple petrophysical model of shale matrix is illustrated in (Ambrose et al., 2010). The
model is used to quantify gas-in-place and it is typically done by methods developed specifically
for tight rocks and other low-permeability formations such as Shales (Luffel et al., 1992; Luffel et
al., 1993; and GRI, 1997). However, the effective pore volume is not directly determined in
these studies; rather total porosity, total water volume, and total oil volume (by weight
difference and an assumed oil density of 0.8 g/cc) are determined. Nonetheless, as shown here
in equations (1) and (2), the total and effective gas porosity values are equivalent.
..... (1)
( )( )
b g v g b v g
J J J J J J S = =
3

...... (2)
The bulk volume, V
b
, in equations (1) and (2) is determined by mercury displacement of
competent cores. Grain volume, on the other hand, is determined on crushed cores via helium
porosimetry. The difference between these two volumes yields the total void volume, V
v
,
(associated with the total porosity ) available for all in-situ fluids, i.e., mobile hydrocarbons,
free and bound water, adsorbed gas, solution gas, and free gas.
For total gas storage, the shale gas-in-place volumes are generally considered in terms
of the following:
1. A volumetric component, G
f
, involving hydrocarbons stored in the pore space as
free gas. The free gas volume is quantified by modifications of standard
reservoir evaluation methods.
2. A surface component, G
a
, with the gas physically adsorbed on large surface area
of the micro- and mesopores (see definition of pore size classification in Chapter
2). The adsorbed gas amount has generally been quantified from the sorption
isotherm measurements by establishing an equilibrium adsorption isotherm
(Mavor et al., 2004). Note that there is no direct measurement of adsorbed gas
phase. It is estimated indirectly from the measured temperature and pressure
conditions.
3. A volumetric component, G
so
, involving gas dissolved into the liquid
hydrocarbon. This volume is usually combined with adsorbed gas capacity in
reservoirs that contain a large fraction of liquid hydrocarbon in the pore space.
4. A volumetric component, G
sw
, involving gas dissolved into the formation water.
The amount of dissolved gas is estimated from the bulk solubility calculations.
S
ge
= J
ve
J
b
( ) J
g
J
ve ( )
= J
g
J
b
4

Although it has traditionally not been considered important, a recent study is
available discussing significant enhancement in gas solubility in formation
liquids when confined to small pores (Diaz-Campos et al., 2009).
Hence, we have G
st
as the total gas-in-place, as the sum of all of these volumetric
contributions:
..... (3)
where the components of storage on the right-hand-side are defined as:


........ (4)


In the current industry standard calculations, the Gso and Gsw terms are not applied.
The solution gas in mobile hydrocarbons and water, and the adsorbed gas within organic matter
are combined in the adsorption isotherm analysis; therefore equation (3) is reduced to:
...... (5)
The current volumetric approaches for shale gas are built on the premise that the two
volumes on the right hand side of equation (5), being associated with the inorganic pores and
organic solid, respectively, can be estimated independently of one-another. Accordingly, the
adsorbed gas is associated with the organics, the pore volume of which is considered to be
G
st
= G
f
+ G
a
( )
w
sw w
sw
o
so o
so
L
sL a
g b
o w
f
B
R S
G
B
R S
G
p p
p
G G
B
S S
G

6146 . 5
0368 . 32
6146 . 5
0368 . 32
1
0368 . 32
=
=
+
=

=
sw so a f st
G G G G G + + + =
5

negligible and, therefore, the volume does not need to be accounted for during the calculations
of free gas; whereas, all of the free gas is associated with the inorganic voids such as
macropores, fissures, fractures etc. Recently Ambrose has proposed a new petrophysical model
for Shale Gas in Place calculations (Ambrose et al., 2010), specifically includes a change in G
f

where the density of adsorbed gas in the organic pore,
s
, must be known, equation (6).
...... (6)
Direct measurements of the adsorbed phase density
s
at pressure and temperature
conditions corresponding to a shale gas reservoir was not carried out yet. Previous research
done by Mavor et al point out that the density is either assumed equal to the liquid density at
1atm and -161.6
o
C, this is 0.4223 g/cc (Mavor et al., 2004); or 0.375g/cc which corresponds to
the van der Waals equation of state co-volume constant. There is also a recommendation of
using Kobayashi values. Haydel and Kobayashi measured the adsorbed methane density,
0.374g/cc, however, their values were encountered using an inorganic adsorbent (silica gel) and
at low pressures and temperature conditions, 1000psi and 104
o
F. In this research a density
values within the range 0.28-0.3g/cc was found using Molecular Simulation. It is found that the
value is sensitive to changes in temperature, pressure and pore size. Typically, at 3,043 psi and
176
o
F, the adsorbed phase density of methane is 0.3 g/cc for a pore width of 2.3nm.
Using scanning electron microscopy, it has been repeatedly shown that a significant
portion of the total pore volume is associated with pores in the organics (Wang and Reed, 2009;
Loucks et al., 2009; Sondergeld et al., 2010). Hence, the organic pores contribute significantly to
the total free gas pore volume. Although it might appear as a rudimentary discussion in shale
petrophysics, the observation is important to the gas-in-place calculations. The organic pores
( )
(
(

|
|

\
|
+

=

L
sL
s b
w
g
f
p p
p
G
M S
B
G

`
10 318 . 1 1 0368 . 32
6
6

are typically in a micro size level, thus a portion of the total pore volume could be taken up by
finite-size adsorption layer and not available for the free gas molecules.
In addition, in a porous medium characterized by the presence of small pores and
occurrences of formation water (e.g., free and inherent moisture), such as a coal and shale, gas
can be stored as dissolved gas in the interstitial water. For gas in the large pores, it is known that
the dissolved gas is only a small fraction of the amount of gas pore could hold. The amount of
dissolved gas in the water could increase in smaller pores, however. The amount dissolved then
may represent a significant portion of the total gas reserves in unconventional reservoirs. The
latter argument suggests that we should consider gas solubility in water confined to micropores.
Best quantification of this solubility and insight into this solution chemistry problem ideally
comes from theoretical studies at the pore-scale, particularly from molecular-level simulations
and the subsequent free energy change calculations accompanying dissolution. In this thesis,
natural gas solubility enhancement in water residing in a graphene slit-pore with a length-scale
on the order of nanometers is investigated using molecular dynamics simulations. The solubility
of methane in water is estimated under controlled temperature condition using the test particle
insertion method with the excluded volume map sampling (Stapleton and Panagiotopoulus,
1990). The results indicate that the methane solubility in nano-scale pores is increased
significantly (typically, one order of magnitude larger), compared with that in the bulk phase,
i.e., in the absence of pore walls. The enhancement is further investigated under varying force-
field conditions due to pore-wall surface potential, namely hydrophobicity, and it is found that
the solubility is extremely sensitive to pore wall wettability. The work is a fundamental approach
to better understand the unconventional gas reservoirs, and important for the estimation of gas
reserves.
7

1.2 Purpose of study
It is not possible with the current technology to carry out direct experimental measurement and
analysis of the gas behavior and sorption in small pores under extreme conditions of
temperature, pressure, for example at typical Barnett initial conditions of 4000 psi and 176
o
F.
Computer simulations, however, provide a direct connection between the microscopic
descriptions of the system (potential interaction between molecules, molecular mass, charges,
etc.) and macroscopic properties of interest, such as adsorption profile, solubility, and vapor-
liquid coexistence of hydrocarbons etc.
In this work, based on computer simulation analysis, we argue that the total gas storage
capacity measurements and the resulting gas-in-place values are inherently uncertain. The
source of the uncertainty involves the improper accounting of the pore volumes occupied by the
adsorbed and free gas phases. The work investigates the nature of the uncertainty using
Equilibrium Molecular Dynamic (EMD) and Monte Carlo simulation (MC) of methane, carbon
dioxide and water in pores with sizes varying between 1-4 nanometers (nm).
8


CHAPTER 2
MOLECULAR MODELING AND SIMULATION

Molecular Dynamics and Monte Carlo Simulations have made enormous strides in recent years
and are gradually becoming a common tool in science and engineering. This is due to the
development of new methods for description of complex fluids and materials, and to the
availability of high-speed and large-capacity computers. Today, molecular models of, for
example, multi-component mixtures are common (Nath et al., 1998), and predictions of
structural, mechanical and electronic properties for large molecules are considered to be
reliable (Karayiannis et al., 2002). Furthermore, molecular simulations are being used more
frequently to construct virtual experiments in cases where controlled laboratory experiments
are difficult and too costly, if not impossible, to perform. Molecular investigations of the phase
transformation of materials under extreme pressures are good examples of the latter
application (Lacks, 2000).
The main objective in this study is to predict accurately the thermodynamic state of
natural gas confined to micro- and nanopores using molecular simulation techniques. The
output obtained from a molecular simulation is then used, with appropriate time correlation
functions, to calculate certain macroscopic quantities of interest such as simple thermodynamic
averages (e.g. kinetic, potential and internal energies, temperature and pressure). The most
important macroscopic quantity during this investigation is mass density, to determine the
adsorption profile of methane in a modeled organic wall, and the Ostwald coefficient which can
9

be related to Henrys law constant of solubility. A detailed description of the numerical
approach is summarized below.
Regarding the sorption processes in small pores, IUPAC introduced a classification of
pores as follows (E. Robens et al., 2005). The pores of internal width greater than 50 nm are
classified as macropores, less than 2 nm as micropores, and between 2 and 50 nm as
mesopores. Since the term micro is misleading for the petroleum community, we are going to
refer to the pores less than 2nm as nanopores, to those with sizes in between 2-5nm as
micropores.
2.1 Molecular Dynamics Simulation
Molecular Dynamics (MD) simulation theories lie in classical mechanics, meaning that every
system in nature is in constant evolution. Of all the physics laws, only those which define a
trajectory of a system along the time, without losing information from the past, are the
accepted ones to describe its molecular dynamics.
One of the most important laws in physics is that the total energy and the total
momentum of a system is conserved, i.e., JIJt + JuJt = u or JJt = u and, JpJt = u,
respectively. Total energy, a sum of kinetic plus potential energies (E = I + u), and total
momentum are conserved as the consequence of Newtons law. Hence, MD is just the numeric
solution of the classical Newtons equation.
MD uses the basic principles of Newtonian mechanics to study thermodynamic and
transport properties of a many-particle system. A system could be a free fluid (bulk) or solid
material, or fluid confined, fluid in a pore or porous medium. The system is modeled as particles
(atoms or molecules) with an initial and instantaneously-predicted position and momentum. The
10

particles are allowed to interact for long periods of time to allow the system to be ergodic, i.e., it
is reached the time when it holds no other conserved quantity besides energy. Hence, the
system reaches the equilibrium.
Once the equilibrium has been reached, given a time interval, the entire motion of the
particles in space and time can be defined and it satisfies the deterministic law of physics.
Deterministic laws mean if a property is known at one instant of time in the phase space of the
system, then it is known at any time in future. It is completely deterministic in the sense that not
only the property can be predicted into the future but also into the past.
The trajectory of each particle in space and time represents the evolution, the dynamics,
of the system. Once the trajectory of a system is defined, the information in the microscopic
level is converted into macroscopic terms such as virial pressure, density, internal energy, etc.
All these thermodynamic properties, averaged over a long finite period of time, are derived
from the equations of state and the fundamental thermodynamic equations, taking into account
that a system is constrained to a conserved quantity. Such quantities are the energy (E), density
(), number of particles (N), temperature (T), pressure (p) and chemical potential (). Notice
that the most important conserved quantity is energy. The rule is to study the system subject to
the constraint of some conserved quantity by fixing it to a known value. The value of the
conserved quantity can be taken from any experimental result or a previous computer
simulation work. This means for example we need to know what is the bulk methane density at
4000psi and 176
o
F, we can take that value from a previous simulation or a previous
experimental work. Knowing the density of the system we initialize positioning each particle in a
known volume. Temperature and pressure are fixed and now we run it to find out what it is the
chemical potential. Hence we start the simulation with a bulk density profile and run the system
11

until it reaches the equilibrium. Once the equilibrium has been reached we know now the
chemical potential and that value will be used for the next simulation when the bulk system is
now confined in a porous system.
A collection of various conserved quantities, usually three of them, makes up an
ensemble. If an ensemble is defined by NPT it is referred as an isobaric-isothermal ensemble. If
it is defined by NVE, micro-canonical; NVT, canonical and finally VT is referred to as the grand-
canonical ensemble. The latter ensemble is used in MC simulation.
Since numerical integration of the Newtons equations of motion make up the core of
the simulation, special algorithms (e.g., Verlet, Leap Frog, Beeman algorithms) have been
developed and commonly utilized during a molecular simulation study. Others exist (Smith, et al.
2008) like the deflection algorithm (Ercolessi, 1997). In the present study both Verlet and Leap
Frog algorithms have been used to estimate and analyze the fluid density under typical
reservoir pressure and temperature conditions: (i) runs involving bulk-phase (i.e., in the absence
pore-walls) methane density measurements using a fixed number of methane molecules at fixed
pressure and temperature (NPT), and (ii) runs involving measurements of density of methane at
the same temperature but confined to a small pore with organic (graphite) walls using a
canonical ensemble (NVT); (3) runs involving methane solubility in water confined to small pore
with graphite walls by mean of a canonical ensemble as well.
2.2 Monte Carlo Simulation
There exists an exhaustive literature studying the equilibrium thermodynamics of fluids using
MC simulation involving phase change of bulk fluids (Harris, 1995), characterization of porous
materials using gas adsorption (Aukett, 1992; Sweatman and Quirke, 2001), and multi-
component gas separation (Ghoufi, 2009). Monte Carlo (MC) simulation is basically a stochastic
12

method for sampling points in multidimensional space, according to a prescribed probability
distribution defined on that space. It allows us predict thermodynamic properties such as: Gibbs
free energy (specifically in grand canonical MC), enthalpy, virial pressure, compressibility gas
factor, particle density, etc.; based on the principles of statistical mechanics. There is a
trajectory in phase space generated by MC which samples from a chosen statistical ensemble.
Note that this trajectory is not a dynamic trajectory, there is no time involved in the calculation.
It is solely a trajectory along the phase space. Since in MC it is not important the dynamics of the
system, it is a simpler technique, less computer intensive hence faster technique than MD. The
Monte Carlo method enables the use of certain ensembles specifically designed for computing
phase equilibria (particularly the Gibbs ensemble) that are very difficult to simulate using
molecular dynamics (Marcus Martin, 2008).
As in the case of MD, there are various algorithms used to perform MC simulation. One
of the most useful is the Metropolis algorithm (uses Markov chain) in which first it has to be
designed some rate at which it has to be chosen new states j from an initial state i and then it
should have an acceptance criteria. A typical acceptance criterion is that when the new state has
lower energy than the previous one, then the new state is accepted at 100%. If the new state
has a higher energy than the previous state then the acceptance of the new state is related with
a certain thermal-like probability (Boltzmann factor), such that the detailed balance is
satisfied. Detailed balance states that the rate of going out of a state i into a state j divided over
the rate of going out of a state j into a state i equal the probability of being in a state j over the
probability of being in a state i. If the detailed balance is satisfied then we end up with a steady
state distribution; hence the probability distribution does not change anymore. The multi-
dimensional space sampled is the configuration space spanned by the coordinates of the
molecules constituting the system. The probability density is set by an equilibrium ensemble,
13

i.e., collection of a large number of replications of the system. Sampled configuration-space
points (or states) form a Markov chain (Metropolis algorithm), with each state being formed
from the previous one in a Monte Carlo (MC) step. Each MC step is executed in two stages. The
first stage attempts an elementary move from the current state i into a new state j with
probability (ij), where the stochastic matrix of attempt probabilities is symmetric. The
second stage then accepts (or rejects) the attempted move with a certain probability which is a
function of the probability density set by the equilibrium ensemble. If the move is rejected, state
i is retained for the current step.
As the consequence of this iterative process, the generated Markov chain of states
asymptotically samples the probability distribution; thus, all the configuration-dependent
properties fluctuate around constant average values representative of thermal-equilibrium. It is
said that the equilibrium is reached when the variance in the estimate of the average ensemble
property, <
NVT
>, say in NVT ensemble; is minimized
2
<
NVT
> (Allen and Tildesley, 2007). In
other words the Markov chain has converged. The system will continue in such way that the
energy will always fluctuate around the minimum value. Those states with low energy are the
ones that are sampled in order to find the average thermodynamic quantities.
In this study the Grand Canonical Monte Carlo (GCMC) simulations are used to
investigate the fluid density profile, when the fluid, methane, is confined in small organic pores
under typical shale reservoir pressure and temperature conditions.
2.3 Numerical Procedures for Gas Adsorption
2.3.1 Molecular Dynamics Approach
For the molecular-level investigation, methane, which is the main component in natural
gas reservoirs, is considered at some supercritical condition under thermodynamic equilibrium
14

in three-dimensional periodic orthorhombic pore geometry consisting of upper and lower pore-
walls made of graphene (carbon) layers, see Fig. 2.1. For simplicity it is considered the shape of a
pore as a slit-like pore.
Using MD, for comparison, we take into account three slit-like pores with a pore width
equal to H=3.9 nm (micro pore) and H=2.3 nm (micropore) and 1.1nm (nanopore), respectively.

Figure 2.1 Molecular simulation cell consisting of graphite walls and OPLS-UA methane
Pore width is an important length-scale of the study, which is defined as the distance
between the inner most graphene planes. The pores maintain fixed dimension in the y-
coordinate: L
y
=3.93nm; however, the dimension is changed in the x-coordinate such that both
pores host approx the same number of methane molecules (400-450) during the simulations
and save some computational time as well. Hence, the dimensions in the x direction become L
x

=4.26 nm and L
x
=7.67 nm for the large and small pores, respectively. The fluid density profile
was calculated for the 2.3 nm pore. A total of 400-600 methane molecules are typically used
during the simulations. 100 up to 144 processors have been used for the simulation and the
z
y
x
15

approximate parallel computation time was typically between 48 min and 25 seconds for the
corresponding number of processors respectively.
A united-atom carbon-centered Lennard-Jones potential (based on the optimized
potentials for liquid simulations, OPLS-UA force field), is used as the model of a methane
molecule, Table 2.1 shows the energy and distance Lennard Jones parameters used for fluid-
fluid and solid-solid interactions (Nath et al. 1998). Notice that has energy units (Temperature
multiplied by the Boltzman constant, k
B
, has energy units)
The methane-solid and methane-methane interactions are of the Lennard-Jones (LJ)
type (Frenkel et al., 2002). A Lorentz-Berthelot mixing rule [
ij
=(
ii
+
jj
)/2 and
ij
=(
ii
jj)
1/2
] is set
up in order to describe solid-fluid (i.e., pore wall-methane) interactions. Here,
ij
and
ij
are the
Lennard-Jones (LJ) parameters accounting for interactions between a molecular site of methane
species and a carbon atom of the organic wall. LJ were cut off at 4.1 for the 3.9 pore width, at
3 for the 1.1 and 2.3 nm pore width respectively. and are Lennard Jones parameter where
, in units of energy (kJ/mol or K), and in units of distance () are the depth of the potential
well and the distance between the two molecules when its potential energy is cero.
The molecular simulation package, DL-POLY (version 2.20), was used to perform the
molecular dynamics simulations (Todorov and Smit, 2008). Methane bulk density computations
are done considering isobaric and isothermic conditions with a constant number of atoms,
pressure and temperature (NPT) at three constant temperatures (T), 176
o
F, 212
o
F, and 266
o
F,
using Nos-Hoover thermostat and barostat ensemble (Frenkel and Smit, 2002). In order to
keep pressure and temperature constant it is used an algorithm which will measure those
parameters at a certain interval of time, this time is known as relaxation time, which is less than
16

the time step in the simulation. The relaxation time used in the ensembles was between 8 to12
picoseconds (ps) for the thermostat and 10 to 15 ps for the barostat.
In the case of simulations involving methane in carbon slit-pore, initially the fluid system
was equilibrated at a constant temperature, with a canonical ensemble (constant number of
atoms, constant volume and constant temperature, i.e., NVT) using Berendsen thermostat
algorithm (Frenkel and Smit, 2002). The equilibrium is assumed to be achieved if no drift as a
function of time was observed in the time-independent quantity, such as the total energy of the
system. After the equilibrium is reached, a new simulation is carried out in the canonical (NVT)
ensemble using the Leapfrog algorithm (Frenkel and Smit, 2002; Allen and Tildesley, 2007) and
Nos-Hoover thermostat. Although the Berendsen algorithm smoothly forces the fluid system to
reach certain positions and velocity in which the temperature fluctuation is minimized, it is time-
irreversible and the computations are not considered in the canonical ensemble; the Nos-
Hoover algorithm needs to be used to control the system temperature while the real canonical
ensemble computations are carried out for the density computations (Frenkel and Smit, 2002).
The relaxation time used for the Nos-Hoover thermostat is optimized to 13, 15 and 20 ps for
the pore widths of 1.1, 2.3 and 3.9 nm respectively. The values of the relaxation time needed to
be optimized because very small values are known to cause high frequency temperature
oscillations on the thermostat, whereas values that are too large lead to drift in temperature
(Hnenberger, 2005). Typically, the total run time for a simulation was from 1.2 to 1.5
nanoseconds for the nano and micropores respectively and the time step used was 0.003 ps.
At the end of each simulation run, number density
Number
(number of molecules per a
certain volume, in this case the volume is L
x
=76.7 , L
y
=30.3 and L
z
=0.2) for methane across
the pore space was computed. Notice that
Number
is computed at every Az =0.2 interval (for
17

the continuous density profile) and
CH4
(discrete density profile) at every volume segment
L
z
=3.8 in the z-direction. 3.8 corresponds to the diameter of methane particle. The number
density for each volume segment is estimated by summing the number of molecules counted in
each z-interval within the volume segment and dividing the total by the number of intervals.
Following, the number density is converted to the local mass density of methane using the
molecular weight of methane M
CH4
and Avogadros number as follows:
... (7)
As explained earlier, it is expected that the local mass density for methane across the
pore will be different from its mean bulk density due to changing levels of interactions between
the methane-methane and methane-carbon bodies. The purpose of our numerical investigation
was to predict a precise density profile across the pore as an indication of the presence of these
interactions and to determine an average adsorbed gas density value.



Table 2.1- Lennard-Jones Potential Parameters for Methane-Graphite System.
2.3.2 Monte Carlo Approach
NPT-Gibbs Ensemble Monte Carlo (NPT-GEMC) simulations are used in this work. The
Gibbs ensemble is defined by introducing a system in which the temperature, T, pressure, p, and
number of fluid particles, N are fixed a priori. T = 176 F, p = 3043 psi and N = 2000. This system is
divided into two (separate) sub-systems labeled 1 and 2. The (variable) volumes of these sub-
23
10 02252 . 6 /
4 4
=
CH Number CH
M
atom (nm) /k
B
(K)
carbon 0.340 28.0
methane 0.373 147.9
18

systems are V
1
and V
2
and (variable) number of particles N
1
and N
2
respectively. Note that N = N
1

+ N
2
. System 1 contains the porous medium; system 2 contains methane in ideal gas phase.
System 2 behaves as a reservoir source. During the adsorption process carried out within the
porous medium. The thermodynamic equilibrium is achieved when the Gibbs free energy (G)
reaches a minimum value and fluctuates around that minimum point (Gibbs free energy is kept
constant). This means that the chemical potentials in the adsorbed phase (system 1) and in the
bulk phase (system 2) are equal, which allows one to directly relate the adsorption information
to the bulk phase properties. The simulations were performed in cycles, each cycle having four
steps: a biased volume displacement step, attempt to transfer molecules between the 2 sub
systems, reinsertion moves, translation moves and attempt to rotate a molecule (adsorbate) in
both systems. Notice that since methane is modeled as a neutral sphere LJ particle, the rotation
of the same will not produce any different result than when it is not rotated.
In the volume displacement step system 1, which contains the porous medium is
constrained to have zero displacement probability and system 2 will have a random volume
displacement of maximum 0.1
3
. The maximum displacement was chosen in such a way that the
acceptance ratio was approximately 50 per cent. This is a biased displacement because we want
to avoid change in volume of the porous medium. The next step was to transfer molecules
between both sub-systems, at initial state there is no adsorbate in system 1 and the probability
of transferring it from system 2, is set up to 1. Once system 1 accumulates a certain number of
adsorbates then the probability of placing one particle from one to another system is set up to
5%. The third step is the re-insertion move, set up to 10%; this move takes a molecule out of
one system and tries to place it back into the same system. The last step is the attempt of
performing a center-of-mass translation move on a molecule without regard to which system
the molecule is currently located in. This move chooses a molecule of the appropriate type at
19

random, chooses a vector on a unit sphere at random, and then attempts to displace the entire
molecule a random distance. The target acceptance rate for the center-of-mass translation
move was 50%.
Three pore width systems were modeled, in order to compare it with the results
obtained with Molecular Dynamics. In this case it was elaborated a pore width of 3.8 nm which
differs a little with the pore in MD, however it was considered better the 3.8nm because it will
contain along the width 10 methane molecules and the analysis of the density profile was
simpler. The Lennard-Jones potential was truncated at 4.5
CH4
for the 3.8nm width pore, 3
CH4

for the 2.31nm pore width and at ~1.9
CH4
for the 1.14nm pore width. For each simulation we
performed 20000 cycles.
At the end of each simulation run, number density of methane is calculated from the
last configuration space obtained in system 1. It is counted, on z direction (where the pore width
extends to), the number of particles encountered at every z=3.8 and divided this amount by
the volume in which the correspondingly number of methane fall within. Once the particle
density is known,
Number
, local mass density is calculated using methane molecular weight M
CH4

and Avogadros number as it is shown in equation 7.
The purpose of our numerical MC investigation was to confirm the values of adsorbed
density phase for methane obtained by MD. In that way we can validate both of our results,
since there is no experimental value at the moment to compare with.
2.4 Numerical Procedure for Methane Solubility in Nanopores
The main objective is to accurately predict the solubility of methane in bulk water and in water
confined to nanopore using Test Particle Method (Widom, 1963) combined with the Excluded
20

Volume Map Sampling (EVMS). The output obtained from molecular simulation is then used,
with appropriate time correlation functions, to calculate certain macroscopic quantities of
interest such as simple thermodynamic averages (e.g. kinetic, potential and internal energies,
temperature and pressure) and transport coefficients related to the dynamics of fluid (e.g. shear
and bulk viscosities, diffusive heat and mass transfer coefficients). In our case this macroscopic
quantity is the Ostwald coefficient which can be related to Henrys law constant of solubility. A
detailed description of the numerical method can be found in classical books by Allen and
Tildesley (1987), and Frenkel and Smit (2002).
2.4.1 Fluid Model and Potential Function.
Simulations of fluid in the canonical (NVT) ensemble are carried out using the Leapfrog
algorithm and Nos-Hoover Thermostat at 75.5
o
C.Typically, the bulk fluid system is composed of
891 water molecules in a 3-D parallelepiped computation cell of 3.85x3.46x2.08 nm furnished
with periodic boundary conditions, resulting in 0.961g/cm
3
water density at 75.5
o
C.
There are a number of atomistic models for water (Chaplin, 2004). The so-called
extended simple-point charge water model (SPC/E) characterized by three point masses with a
O-H bond of 0.1 nm, and H-O-H angle equal to 109.47
o
was used in this study. The centered
electron charges on the oxygen and hydrogen are -0.8476e and +0.4238e, respectively. O-H
bond is kept rigid during the simulations using SHAKE algorithm. A time step of 0.2 fs was
assigned to maintain good energy conservation.
Lennard-Jones interaction is used to represent the van der Waals forces between the
oxygen-oxygen atoms of the two SPC/E water:
u(r
]
) = 4e _[
c

12
- [
c

6
_. (8)
21

where = u.6Su4 k}mol and o = S.16S6 are the depth of the potential well and the
distance between the two water molecules respectively (when its potential energy is at
minimum); r is the inter-particle distance at real time during the simulation. Calculations of long-
range water interactions were performed using Ewald sum technique. The total interaction
energy between the water molecules consists of the Lennard-Jones potential and the Coulombic
potential based on the classical electrostatics: U
TOTAL
= U
LJ
+U
C
, where U
C
between two molecules
i and j is represented as the sum of Coulomb interactions acting among the charged points in
the following way:
u
C
=
q
i
q
]

i]
L] 0,H
.. (9)
where r
]
L]
is the distance between site L of the molecule i and site J of the molecule j.
A united-atom carbon-centered Lennard-Jones potential (based on the optimized
potentials for liquid simulations, OPLS-UA) force field, =1.2309 kJ/mol and =3.73 , has been
used in the simulations of methane. Methane is considered as a spherical molecule and the
water-methane interaction was described by the optimized Konrad-Lankau (2005) potential, see
Table 2.2, the methane-methane interaction is not required since methane is the test particle.
Fluid-fluid and fluid-solid interactions are also of the Lennard Jones type and act between
methane-water, water-water, graphite-water and graphite-methane respectively. The LJ
interactions were cut off at 0.9nm.
2.4.2 Small Pore Model.
Graphite is modeled as a slit-like pore, using the Lennard Jones conventional potential
where each plane of the graphitic surface, i.e., graphene, consists of stacked planes of carbon
atoms separated by 3.4 distance. The carbon-carbon separation on the plane along a hexagon
22

side is 0.142nm.The distance between two graphene layers which confine the fluids is
considered as the width of the slit-pore system. The solubility study is numerically performed at
four separate confined systems with pore width values of 6.9, 11.6, 16.2 and 20.8. As the pore
width is increased, the number N of water molecules varies from 297 until 891. Standard
Lorentz-Berthelot combination rule is used to obtain the corresponding LJ parameters and
describe the interactions of graphite-methane and graphite-water:
o
]
=
1
2
|o

+o
]]
] and e
]
= _(e

e
]]
) ..... (10)
where , and
jj
,
jj
correspond to the distance and energy parameters of the pure
chemical species case. The distance and energy parameter for describing graphite-methane
interaction are modified by a constant of 0.97 and 1.66 respectively. The total length of each
MD run was 400 picoseconds.
interacting molecules (kJ/mol) ()
SPC/E water- SPC/E water 0.6504 3.1656
OPLS-UA methane-SPC/E water 1.0131 3.56
graphite-graphite 0.2353 3.4
graphite-OPLS-UA methane 0.8947 3.4478
graphite-SPC/E water 0.3912 3.2828

Table 2.2- Lennard-Jones potential parameters for methane-water-graphite system.
2.4.3 Infinite Methane Dilution and Widom Test Particle with EVMS.
Usually the solubility of a solute (2) in a liquid solvent (1) is measured by the Ostwald
coefficient, L =
2
I

2
g
where
2
I
and
2
g
are the number densities of the solute in the liquid and
gas phases in equilibrium, respectively. The condition of equilibrium provides another
expression for L, which is:
23

L = e
-(
2
-I
-
2
-g
)k
E
T
..... (11)
where
2
-I
and
2
-g
are the excess chemical potentials of the solute in the liquid and gas phases,
respectively. Hence, in general, the evaluation of the solubility requires knowledge of the excess
chemical potential of the solute in both phases.
When the system density in the gas phase is negligible, Ostwald coefficient becomes
identical to the liquid phase activity coefficient
I
and the only required quantity for its
evaluation is the excess chemical potential of the solute in the liquid phase alone (Guillot and
Guissani, 1993):
L
I
= e
-
2
-I
k
E
T
...... (12)
In statistical mechanics the excess chemical potential
-
of one molecule of solute in a fluid
solvent, i.e., the infinite dilution limit, can be given in terms of the Helmholtz free energy
difference between the system composed of (N+1) molecules and the one composed of N
solvent molecules:

-
= F = -Ik
B
ln (
z
T,N+1
vz
T,N
) ... (13)
Here, Z
T,N
is the configurational partition function at temperature T and volume V. Using test
particle method described by Widom (1963) Eqn. 10 can be re-written as

-
= -k
B
Iln(e
-
k
B
T
)
N
= -k
B
Iy ..... (14)
where is the potential energy difference between a system composed of (N+1) molecules and
that containing only the N solvent molecules. Consequently, it is possible in practice to
24

introduce methane, as a test particle, at a fixed vacant location into the solvent and evaluate the
Boltzman factor exp
(-/kBT)
, while moving the whole system using the MD simulation.
In addition to the Ostwald coefficient, the solubility of gases in liquids can be expressed
in terms of Henrys law constant, k
H
. In the liquid phase k
H
is simply related to the activity
coefficient
l
with the expression:
l
H
RT
k

=
......... (15)
where is the number density of the pure solvent.
Before calculating the potential energy of the interaction between the test particle that
has to be inserted and the solvent, a map is made of the previous solvent system in equilibrium,
this map is done by dividing the system into small boxes that can contain at most one molecule
of solvent. Each box is labeled depending on whether it contains a particle or it is empty. This
map is used to check whether there is an empty cavity or space for the particle to be inserted. In
our case, we divided each configuration into a series of cubic sub-cells with length of 3.1656
and inserted in an empty cavity the methane molecule. See Figure 2.2.
Prior to the insertion of methane, the system was equilibrated at 75.5
o
C using Berendsen
thermostat algorithm. After the equilibrium has reached, Nos-Hoover thermostat algorithm is
used to control the system temperature and carry out the real canonical ensemble on the N and
N+ 1 particle fluid systems. The equilibrium is assumed to be achieved if no drift was observed
in the time-independent quantity, such as the total energy of the system. Figure 2.3 shows the
total energy of water system under equilibrium conditions in a nano slit-pore at 75.5
o
C.
25

At frequent intervals during the simulation we generate, by using the excluded map
sampling method, a coordinate with N+1 particles, we then compute the Boltzmann factor exp(-
/kBT), sampling the last 360 canonical configurations for each confined system. The interval
for the insertion was every 50ps. In the case of the bulk model a trajectory of 120ps was
sampled for analyzing the Boltzmann factor.

Figure 2.2 Methane insertion at a vacant spot.

Figure 2.3 Total energy (left) and temperature (right) fluctuation over the time for a pure water
system in thermodynamic equilibrium in a nanopore of 1.5nm at 75.5
o
C
2.4.4 Pore Walls with Controlled Wettability.
When a small amount of two immiscible fluids are put in contact with a flat solid
surface, three phases are delimited by a certain triple contact line. Since each interface has its
y = 7E-05x - 31287
-31,294
-31,292
-31,290
-31,288
-31,286
-31,284
-31,282
-31,280
0 100 200 300 400
TOTAL ENERGY EVALUATION
Time, ps
0.2fs timestep, 348.7K, Nos-Hoover thermostat
y = 5E-06x + 348.7
348.0
348.2
348.4
348.6
348.8
349.0
349.2
349.4
0 100 200 300 400 500 600
TEMPERATURE EVALUATION
Time, ps
0.2fs timestep, 348.7K, Nos-Hoover thermostat
T
e
m
p
e
r
a
t
u
r
e
,
K
26

own free energy per unit area, the fluid-solid interaction energy will determine whether the
solid surface is non-, partial- or totally wet by the correspondent fluid. The partial or total fluid-
solid wetting can be defined as surface hydrophilicity/hydrophobicity depending on whether the
fluid is water or methane. For low-energy surfaces such as graphite, its hydrophobicity can be
driven by a strong methane-wall interaction as well as by a weak water-wall interaction,
promoting an easier sliding for the water particles across the solid (Roselman, 1976).
In principle a given surface is wetted completely if the solid-fluid attraction (i.e., van der
Waals interaction near the surface) is much higher than the fluid-fluid attraction. Using MD
simulation Cao et al. (2006) have earlier shown that a small slit-like pore structure can have an
enhanced surface hydrophobicity due to changes in the liquid-solid interaction. In this study, we
varied the LJ energy parameter for the potential interaction between water-graphite as follows:
graphite water s f
c

=
........................................................................... (16)
where the coefficient c takes values of 0.4, 0.6 and 1.0, in order to see the hydrophobic (when c
= 0.4 and 0.6) of the wall over the solubility.
27


CHAPTER 3
NUMERICAL RESULTS AND DISCUSSION

3.1 Methane Adsorption in Organic Pores using Molecular Simulations.
3.1.1 Free Methane Density Calculations at Reservoir Conditions.
Table 3.1 shows that the predicted free gas density at the center of the pore
corresponds to the bulk methane density predicted independently by National Institute of
Standards and Technology (NIST) thermo-physical properties of hydrocarbon mixtures database,
SUPERTRAPP. The comparison shows that our numerical simulations based on molecular
modeling are accurate.
3.1.2 Methane Adsorption Confined to Micro- and Nano-scale Pores.
Figures 3.1 and 3.2 show density profiles for methane confined to the large pore and the
small pore. It is clear that the predicted methane density is not uniform across the pores: its
value is significantly greater near the wall, where adsorption takes place, and decreases with
damped oscillations as the distance from the pore wall increases. The oscillations are due to
presence of adsorption in the pores and involve structured distribution of molecules, i.e.,
molecular layers. The layers indicate the existence of thermodynamic equilibrium in the pore.
The number of molecules is largest in the first layer near the wall indicating physical adsorption.
The wall effect becomes significantly less in the second layer, indicating that desorption of some
methane molecules is allowed due to equilibrium adsorption. The molecules in the second layer
are still under the influence of pore walls although intermolecular interactions among the
28

methane molecules begin dominating, not allowing locally high methane densities. In this layer,
the density of methane is slightly larger than the bulk gas density of methane (at the center of
the pore). The pore pressures are around 3,043psia which is a quantity predicted using the free
gas molecules and the volume at the center of the pore.
pore
temperature
o
F
pore
pressure
psi
free gas density
at the center of
pore,
g/cm
3

methane bulk
density**,
g/cm
3

176 2206 0.090 0.089
176 3043 0.124 0.122
176 3226 0.133 0.129
176 3676 0.147 0.144
176 3994 0.156 0.154
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 4878 0.181 0.179
176 6272 0.214 0.211
176 7300 0.231 0.229
176 7550 0.235 0.233
176 8707 0.253 0.250
** from NIST-SUPERTRAPP

Table 3.1- Methane density at different pressures. Comparison with the bulk density values
obtained using SUPERTRAPP under the same pressure and temperature conditions.
29

Pore width = 3.93nm




Figure 3.1- Number density (left) and discrete density (right) profiles for methane in 3.9 nm
graphite slit-pore at 176
o
F (80
o
C). Density values are estimated at each 0.2 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.
41
7
5
0
8
16
24
32
40
0.0 3.8 7.6 11.4 15.2 19.0

C
H
4
,

(
1
0
0
0
A
-
3
)
pore half length,
0.281
0.163
0.133
0.126
0.124
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.0 3.8 7.6 11.4 15.2 19.0

C
H
4
,

g
/
c
c

pore half length,

0.331
0.124
0.0
0.1
0.2
0.3
0.4
0.0 3.8 7.6 11.4 15.2 19.0
30

Pore width = 2.31nm


Figure 3.2 Number density (top) and discrete density (bottom) profiles for methane in 2.31 nm
graphite slit-pore at 176
o
F (80
o
C). Density values are estimated at each 0.2 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.
43
8 8
43
0
8
16
24
32
40
48
0.0 3.8 7.6 11.4 15.2 19.0 22.8

C
H
4
,
1
0
0
0

-
3
pore length,
4-5 (bulk)
0.292
0.171
0.135
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.0 3.8 7.6 11.4

C
H
4
,

g
/
c
c

pore half length,

0.350
0.124
0.0
0.1
0.2
0.3
0.4
0.0 3.8 7.6 11.4
31

The observed density profiles show that the assumption of Langmuir adsorption theory
with monolayer is reasonable to describe the equilibrium adsorption dynamics of natural gases
in the organic pores.
3.1.3 Pore Size Effects on Methane Adsorption.
In essence, supercritical methane in small organic pores is structured due to pore wall
effects and shows layers of graded density across the pore. Depending on the pore size, a bulk
fluid region may exist at the central portion of the pore, where the influence of molecular
interactions with the pore walls is very weak. In pores with sizes up to 50 nm a combination of
molecule-molecule and molecule-wall interactions dictates thermodynamic states of the gas and
its mass transport in the pore (Krishna, 2009). On the other hand, within a small pore, see Figure
3.2 and 3.3, methane molecules are always under the influence of the force field exerted by the
walls; consequently, no bulk fluid region can be observed in the pore, hence, no pore pressure
can be measured, therefore, behavior of the adsorbed molecules should be considered rather
than motion of the free gas molecules. Further, the adsorbed-phase density at the first layer is
significantly large value. The smaller the pore the greater the adsorbed phase density. In the
limit, Langmuir theory is not suitable for description of physical interactions between gas and
solid.
3.1.4 Effect of Temperature on Methane Adsorption.
The effect of temperature on methane density is shown in figure 3. 4. The estimated
average adsorbed methane density is around 0.3 g/cm
3
although it decreases with temperature,
for example to 0.259 g/cm
3
at 266
o
F. These values show variations due to changing levels of
kinetic energy at the microscopic scale.
32

Pore width = 1.14nm


Figure 3.3 Number density (left) and discrete density (right) profiles for methane in 1.14 nm
graphite slit-pore at 176
o
F (80
o
C). Density values are estimated at each 0.2 interval for the
continuous density profile. Discrete density corresponds to molecular layer density for methane
across the pore.
47
16
46
0
10
20
30
40
50
0.0 3.8 7.6 11.4

*
C
H
4
,
(
1
0
0
0

-
3
)
pore length,
4-5(bulk)
0.307
0.234
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.0 3.8

C
H
4
,

g
/
c
c

pore half length,
33





Figure 3.4 Number density (top) and discrete density (bottom) profiles for methane in 3.93nm
slit-pore as a function of temperature.
0
8
16
24
32
40
0.0 3.8 7.6 11.4 15.2 19.0

*
C
H
4
,

(
1
0
0
0

A
-
3
)

pore half length,
176 F
212 F
266 F
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.0 3.8 7.6 11.4 15.2 19.0

C
H
4
,
g
/
c
c

pore half length,
176 F
212 F
266 F
34

3.1.5 Methane adsorption confined to nano-scale pores using Monte Carlo Simulations
Gibbs ensemble NPT Monte Carlo results are shown in Figure 3.5. Figure 3.5 confirms
the result obtained by Molecular Dynamic simulation using canonical ensemble. When the
system gets a large pore size the monolayer adsorption phase decreases and there is a uniform
profile along the center of the pore. See dashed distribution along 1.14 and 1.9nm. While in
small pore size widths the uniform behavior at the center of the pore disappears.


Figure 3.5 Density profiles for methane in three different-size slit-pore 1.14, 2.3 and 3.8nm
using MC simulation. Estimated pressure at the center of the pore is 3043psi.
0.300
0.286
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.0 3.8 7.6 11.4 15.2 19.0

C
H
4
,

g
/
c
c

pore half length,
1.14 nm
2.3 nm
3.8 nm
35

Carbon Dioxide


Figure 3.6 Number density (top) and discrete density (bottom) profile for carbon dioxide in
3.8nm slit-pore as function of distance from the wall to the center of the pore, at p = 1980psi.
100
25
15
25
97
0
20
40
60
80
100
0.0 3.2 6.4 9.6 12.8 16.0 19.2 22.4

C
O
2
,
(
1
0
0
0

-
3
)
pore length,
0.910
0.564
0.419
0.372
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
0.0 3.2 6.4 9.6

C
O
2
,


g
/
c
c

pore half length,
36

3.2 Carbon Dioxide adsorption confined to nano-scale pores using Molecular Dynamics Simulations
Molecular modeling and simulations using Carbon Dioxide as supercritical gas is important for
the development of new technologies in subsurface sequestration of greenhouse gases. In this
study carbon dioxide was modeled as a linear rigid triatomic molecule with three charged
Lennard Jones centers having point partial charges at the center of the oxygen (-0.332e) and at
the center of the carbon (+0.664e) to represent the strong quadrupole moment that has CO
2
.
The oxygen-oxygen distance was 0.2324 nm. The pore width modeled for this system was
2.3nm. In total 553 CO
2
molecules were used and the density profile results are shown in figure
3.6. The approximate pressure at the center of the pore is about 1980 psi. It can be observed
that even though the pressure, at the center of the pore, is low; the adsorption selectivity of the
organic wall is 3 to 1 in favor of CO
2
when compared with CH
4
adsorption confined to 3043 psi.
Unless the phase of carbon dioxide at these conditions of pressure and temperature is
uncompressible liquid, it is fairly easy to convince ourselves that at higher pressures, say 3043
psi, CO
2
adsorbed phase should increase at higher values than 0.91g/cc.
3.3 Methane Solubility Enhancement in Water Confined to Nano-scale Pores using Molecular
Dynamics Simulation
For each system modeled, values of the solubility parameters:
l
, L, and the Henrys law
constant, k
H
, of the methane activity in water are estimated at 75.5
o
C and reported in Table 3.2.
The results also include the estimated water density values at the same temperature. At 75.5
o
C
the density of the water vapor coexisting with the liquid phase is considered negligible (Guillot
and Guissani, 1993), therefore methane activity coefficient in the water vapor phase,
g
, is
assumed to be ~1 and the Ostwald Coefficient L ~
l
. The corresponding experimental data for
the bulk (i.e., without pore) is from Economou et al. (1998). It is clear that the numerically-
37

obtained density and k
H
values of this study compare quite well with the experimental
measurement. If the authors of the experimental work were to calculate the liquid phase
activity coefficient l, they would have predicted 25x10
3
which is also a value quite close to that
obtained in this study using MD simulation. Due to the lack of experimental and simulation
results of methane solubility in water under confinement and since the methodology we carried
out for each confined system follows the same process than the bulk one, we consider valid our
results even though they may carry out large statistical errors. Evaluation of Henrys law
constant for methane in water by Economou et al. (1998) illustrates that inserting a molecule by
using the Widom process in an energetically favorable position carries out a larger statistical
error, 10-22%.
Figure 3.7 shows the estimated methane solubility (L ~ ) in water for varying slit-pore
widths and compares with the bulk case. Notice the enhancement in solubility when the fluid
system is confined to smaller pores. When the pore-length is in 1-2 nm range, the estimated
solubility is approximately 2-8 times more. The solubility confined to a slit pore of 0.7nm, a size
comparable to macro-molecular openings in coal and kerogen, is enhanced with a factor of 13.7.
Pore width
()

6.9

11.6

16.2

20.8
Bulk
(numerical)
Bulk*
(empirical)

l
~ L, (x10
3
) 357 54 198 34 135 14 62 13 26 4
k
H
( kbars) 4.3 7.8 11.5 25.1 60.2 63
(g/cm
3
) 0.960 0.960 0.960 0.961 0.961 0.961
N 297 495 693 891 891

Table 3.2- Calculated values of the solubility parameters l, L and Henrys law constant k
H
at
75.5
o
C; N is the number of water particles per each slit pore and bulk system.

38




Figure 3.7 Estimated Ostwald coefficient, L (x10
3
), of methane in water confined to slit-
pores with varying sizes compared with the solubility in bulk water (top). Corresponding SCF
Methane/BBl water (down). Calculated values at 200psi and 168F.
Comparing our simulation results with the study of the temperature dependence of
methane solubility in water (Guillot and Guissani 1993) in a bulk system; methane solubility of a
357
198
135
62
26
0
50
100
150
200
250
300
350
400
6.9 11.6 16.2 20.8 BULK
O
s
t
w
a
l
d

C
o
e
f
f
i
c
i
e
n
t
,

L
(
*
1
0
3
)
Pore width,
Methane Solubility in Water
22.2
12.3
8.4
3.8
1.6
0
5
10
15
20
25
6.9 11.6 16.2 20.8 BULK
S
C
F

M
e
t
h
a
n
e
/
B
B
L

w
a
t
e
r
Pore width,
39

confined system by a slit pore of 0.7nm at 75.5
o
C equals to the bulk methane solubility that
Guillot and Guissani (1993) have found at 336.8
o
C.
Table 3.3 and Figure 3.8 show the effect of varying pore wall wettability on the
solubility. Here, we control the wall wettability by changing the parameter c in equation (16).
When c takes values smaller than 1, we can say that it represents a hydrophobic wall because
the water-wall interaction has been minimized. In the same way when c is larger than 1, we can
say that this represents a hydrophilic wall. When the coefficient c is 0.4 we observe a factor of
1764/135=13.1 increase with respect to the c=1.0 case and a factor of 1764/26=67.9 increase
with respect to the solubility in bulk water; on the other side when c increases up to 1.5, then
methane solubility increases in a factor of 24.3 with respect to when c=1. Indeed, the pore wall
wettability has an influence on the enhancement of methane solubility in water either way it is
hydrophilic or hydrophobic. This enhancement is more dramatic when the wall-water
interaction increases, in other words when the wall is more water wet than methane wet. P. G.
de Gennes in his review of Wetting: Statics and Dynamics, explains that most molecular liquids
achieve complete wetting with higher surface-liquid than liquid-liquid interactions near the
surface (U
Sf
>U
ff
); meaning that if we want to wet completely any surface then the surface-fluid
interaction must be higher than the fluid-fluid interaction. Surface-fluid interaction in our case
corresponds to 0.3912kJmol
-1
while fluid-fluid to 0.6504kJmol
-1
(water-water). An increment in c
yields to a higher surface-water interaction and, therefore we can say that the wall is more
water wet or more hydrophilic. Correlating methane solubility of 1.76x10
-3
g/cm
3
[McCain,
1990], the enhancement produced by confining methane-water solution at 16.2 of slit pore
and having a more hydrophilic pore wall than the organic graphitic one, is two orders of
magnitude (0.22g/cm
3
). The enhancement on methane solubility, when decreasing water
wettability of the wall - having c less than 1, is lower because the interaction difference between
40

methane-methane and methane-wall interaction is higher than water-water and water wall
interaction.

c 0.4 0.6 1.0 1.2 1.5

l
~ L, (x10
3
) 1,764 304 135 2,025 3,283
k
H
( bars) 876 5090 11479 763 471

Table 3.3- Effect on the solubility parameters l, L at 75.5
o
C by changing the wall-water
interaction. The pore width = 1.62nm.



Figure 3.8 Estimated Ostwald coefficient, L (x10
3
), of methane in water confined to 1.62nm
pore with varying pore-wall wettability.
1,764
304
135
2,025
3,283
0
500
1,000
1,500
2,000
2,500
3,000
3,500
0.4 0.6 1.0 1.2 1.5
O
s
t
w
a
l
d

C
o
e
f
f
i
c
i
e
n
t
,

L
(
*
1
0
3
)
c, (
final
= c
initial
)
Wettability effect on Methane Solubility
41

3.4 Importance of Molecular Simulation Findings
In order to show the practical significance of our numerical investigation in shale gas
reservoir engineering, a shale gas in-place study is performed below. The study involves
estimation of free and adsorbed gas volumes of two shales with high and low sorption
capacities. The formula used for the volumes G
f
and G
st
are from Ambrose et al. (2010). The
adsorbed densities chosen are:
1. 0.29g/cc, corresponds to the value predicted in this study at 3043 psi and 176
o
F,
we are considering the average density between two pore widths, larger ones.
2. 0.37g/cc, corresponds to the value estimated by Kobayashi and Haydel and,
3. 0.42g/cc, that corresponds to the liquid methane density, at atmospheric boiling
point condition.
Both shale samples, low and high adsorbed gas volume see table 3.4, considered in
Ambroses et al. research, are taken into account for the comparison.
Shale A:
(low sorption capacity)
Shale B:
(high sorption capacity)
0.06 0.06
0.35 0.35
0.0 0.0
0.0046 0.0046

16 lb/lb-mol 16 lb/lb-mol
50 scf/ton 120 scf/ton

4000 psia 4000 psia
T
180
o
F 180
o
F
p
L
1150 psia 1800 psia
2.5 g/cm
3
2.5 g/cm
3

s1

0.29 g/cm
3
0.29g/cm
3

s2
0.37 g/cm
3
0.37 g/cm
3

s3
0.42 g/cm
3
0.42 g/cm
3


Table 3.4- Shale properties, low and high sorption capacity, taken from (Ambrose et al., 2010),
including three different density phases for comparison.

S
w
S
o
B
g
`
M
G
sL
p

b
42

Free gas storage capacity, G
f
, for Shale A is obtained by replacing
s
for
s1
,

s2
and

s3
.
For example, when the adsorbed density corresponds to
s1
=0.29g/cc; the free and total storage
capacities for shale A are 89scf/ton and 128scf/ton respectively.
scI/ton 89
1150 4000
4000
50
16 10 318 . 1
5 . 2
) 35 . 0 1 ( 06 . 0
0046 . 0
0368 . 32

`
10 318 . 1 ) 1 ( 0368 . 32
6
6
=
(

\
|
+

=
(

|
|

\
|
+

s
L
sL
s b
w
g
f
p p
p
G
M S
B
G


Notice that Langmuir storage capacity, G
sL
, and Langmuir pressure p
L
, are different for
Shale A and B. Hence G
a
for Shale A is 38.8scf/ton.
scI/ton 8 . 38
1150 4000
4000
50 =
+
=
a
G

scI/ton 128 scI/ton 8 . 38 scI/ton 99 = + =
st
G

Using the same procedure and replacing the corresponding parameters for Shale B, its
free and total storage capacity are 66.7scf/ton and 149scf/ton respectively.
Table 3.5 summarizes the obtained G
f
and G
st
values for Shale A and B at different
adsorbed phase densities as mentioned before. When shale gas is of high sorption capacity, the
variance in the estimated G
f
and G
st
values is larger. It is evident that when the inappropriate
s

value is used, the free and total storage capacity calculated for shale A, is increased in 4.2 and
6.1scf/ton. For Shale B it is increased in 9.1scf/ton and 13scf/ton. In percentage, the free storage
capacity is increased in 4% and 7%, for Shale A and in 13.6% and 19% for Shale B. The total
storage capacity is increased in 3% and 4.7%, for Shale A and in 6.7% and 8.7% for Shale B.

43

In our calculation of the gas in-place we predict dramatic change in the estimated gas
in-place values due to volume adjustments. This change in gas in-place persists even when the
sorbed density values reported by other researchers are used, such as the liquid methane
density of 0.4223g/cc corresponds to 1atm and -161
o
C (Mavor et al., 2004) and of 0.374g/cc
(Kobayashi et al., 1967).

Adsorbed Adsorbed Adsorbed Adsorbed
phase phase phase phase
density density density density
lree storage oapaoity lree storage oapaoity lree storage oapaoity lree storage oapaoity 1otal storage oapaoity 1otal storage oapaoity 1otal storage oapaoity 1otal storage oapaoity
Reference g/om g/om g/om g/om
3 33 3
3Cl/ton 3Cl/ton 3Cl/ton 3Cl/ton 3Cl/ton 3Cl/ton 3Cl/ton 3Cl/ton
s Shale A Shale B Shale A Shale B
M3 M3 M3 M3
0.29 89.0 66.7 128 149
aydel & Kobayashi aydel & Kobayashi aydel & Kobayashi aydel & Kobayashi
0.37 93.2 75.8 132 159
Mavor (at 1atm, Mavor (at 1atm, Mavor (at 1atm, Mavor (at 1atm, - -- -
161.2C) 161.2C) 161.2C) 161.2C)
0.42 95.1 79.7 134 162

Table 3.5- Free and total storage capacity for Shale A and B, at different adsorbed density phase
values.


44


CHAPTER 4
CONCLUSIONS AND RECOMMENDATIONS

1.1 Conclusions
The following can be concluded in regards to the gas behavior in organic shales:
1. The adsorbed phase follows Langmuir theory in pores larger than approximately 2 nm and that
it takes up a nearly one-molecule thick portion of a pore, although there is a damped oscillation
density profile with an increased density in the second layer. For a 100 nm pore, the volume is
fairly insignificant; however, for pores on the order of a 10 nm, it is quite large.
2. The shale gas producers should not disregard the volume occupied by the adsorbed phase
during their reserve estimation calculations, because they inadvertently overestimate the pore-
volume available for the free-gas storage.
3. Through MD simulation and Langmuir theory we showed that the density for methane at the
pore wall typically equals to 0.29 g/cm
3
. This value now corresponds to numerical work which
has two limitations: (i) simulations have been done on flat organic pore wall surfaces and in
small pores the adsorbed layer thickness may be affected by the roughness and curvature of the
solid surface, (ii) molecular dynamics simulations in the canonical ensemble in the presence of
multiple fluid phases need to further supported by simulations in the Gibbs ensemble. One way
of approaching the latter problem is using PT Monte Carlo simulations. Investigations using
GCMC shows that the predictions based on MD are reasonable.
45

4. It is found that the level of confinement in pores affects the thermodynamics of fluids
significantly. Consequently, as the pore size reaches nano-scale sorption becomes considerably
larger. The solubility is extremely sensitive to the changes in the small pore wall wettability.
5. By using molecular modeling and simulations it is feasible to measure density of the adsorption
phase, which can be used during gas storage capacity calculation and minimize its uncertainty.
6. Molecular simulations provide the capability to avoid significant errors in gas storage capacity
measurements that results from laboratory methodologies, because of the overwhelming and
expensive procedure to follow up during sample preparation and handling. Such as: minimize
sample exposure to air, light, excessive temperatures; removal of extraneous materials like
drilling fluid; minimize desiccation; samples with inert gas Argon, etc.
7. Carbon dioxide adsorption calculation shows selectivity ranging from 3 to 1 in favor of CO
2
when
compared to methane adsorption at about 1980 psi.
8. Shale gas depleted reservoirs may be suitable for storing carbon dioxide because of the organic
nature of the porous medium and selectivity for adsorbing three times more CO
2
than CH
4
. In
the same context, it can be inferred that methane production can be increased by injecting CO
2
.
9. In our calculation of the gas in-place we predict dramatic change in the estimated gas in-place
values due to volume adjustments. This change in gas in-place persists even when the adsorbed
density values reported by other researchers are used.
10. The enhancement produced by confining methane-water solution at 16.2 hydrophilic slit pore,
is two orders of magnitude (0.22g/cm
3
) compared to the bulk methane solubility of 1.76x10
-3

g/cm
3
.



46

1.2 Recommendations
1. Future work includes re-evaluation of the concepts and approaches that are presented and
discussed in the presence of multi-component gases with varying pore sizes. It is expected that
there will be added uncertainties in the gas-in-place estimation due to (1) multi-component
nature of gas and its adsorbed phase, and (2) volume fraction of organic nanopores that are
currently inaccessible for visual observation and measurement. The former requires MD
simulation studies on the adsorbed phase thickness and density using binary, ternary, and
quaternary mixtures of gases with components such as ethane, propane, and carbon dioxide.
The latter uncertainty, on the other hand, can be reduced by measuring the absolute area of the
organic surfaces using isotherm data.
2. In addition, a new study on transport properties of natural gases in nanoporous organic
materials is necessary using molecular modeling and simulations in the light of the new pore-
scale findings.
3. Solubility studies using hydrocarbon liquids is also recommended for the gas-in-place
calculations.
47


REFERENCES
[1] Ray J. Ambrose, Robert C. Hartman, Mery Diaz-Campos, I. Yucel Akkutlu, and Carl H.
Sondergeld. Shale Gas-in-place Calculations Part I - New Pore-scale Considerations.
SPE 131772, submitted December 2010.
[2] Annual Energy Outlook 2010 with Projections to 2035. Energy Information
Administration 2010, Report Number: DOE/EIA-0383(2010), available at:
http://www.eia.doe.gov/oiaf/aeo/gas.html.
[3] Long, G. U.S. Crude Oil, Natural Gas, and Natural Gas Liquids Reserves. E.I.A. 2009,
available at:
http://www.eia.doe.gov/oil_gas/natural_gas/data_publications/crude_oil_natural_gas_
reserves/cr.html.
[4] Allen, M.P. and Tildesley, D.J. "Computer Simulation of Liquids". Oxford University Press,
London, pp 4-16, 20-28, 33-36, 52-65, 71-108, 110-139, 156-159; 2007.
[5] Cao, B.Y., Chen, M. and Guo, Z. Y. "Liquid flow in surface-nanostructured channels
studied by molecular dynamics simulation", Physical Review E. Vol: 74 pp 66311, 2006.
[6] Chaplin, M. "Water Structure and Science" 2009. available at:
http://www.lsbu.ac.uk/water
[7] Comer, J. B. and Hinch, H. H. "Recognizing and Quantifying Expulsion of Oil from the
Woodford Formation and Age-Equivalent Rocks in Oklahoma and Arkansas", AAPG
Bulletin 71 (7) pp 844-858, 1987.
[8] de Gennes, P.G. " Wetting: Statics and Dynamics" Reviews of Modern Physics Vol: 57 pp
827-863, 1985.
48

[9] Economou, I.G., Boulougouris, G.C., Errington, J.R., Panagiotopoulos, A.Z. and
Theodorou, D.N., " Molecular Simulation of Phase Equilibria for Water-Methane and
Water-Ethane Mixtures" J. Phys. Chem., B. Vol: 102 pp 8865-8873, 1998.
[10]E. Robens, P. Staszczuk, A. Dabrowski and M. Barczak. "The Origin of Nanopores" Journal
of Thermal Analysis and Calorimetry, Vol: 79 pp 499507, 2005.
[11]Ercolessi, F., "To Molecular Dynamics" at http://www.fisica.uniud.it/~ercolessi/md/,
1997.
[12]Frenkel, D. and Smit, B., "Understanding Molecular Simulation: From Algorithms to
Applications", 2nd. ed. (Academic Press, 2001).
[13]Guillot, B. and Guissani, Y. "A Computer Simulation Study of the Temperature
Dependence of the Hydrophobic Hydration" J. Chemical Physics, Vol: 99:10 pp 8075-
8094, 1993.
[14]Haydel, J., and Kobayashi, R. "Adsorption Equilibria in the Methane-Propane-Silica Gel
System at High Pressures" Industrial and Engineering Chemistry Fundamentals, 6: 564-
554, 1967.
[15]Hnenberger, P.H. "Thermostat Algorithms for Molecular Dynamics Simulations". Adv.
Polym. Sci. Vol: 173 pp 105-149, 2005.
[16]Jenden, P. D., Drazan, D. J.,Kaplan, I. R. "Mixing of Thermogenic Natural Gases in
Northern Appalachian Basin", AAPG Bulletin 77 (6) pp 980, 1993.
[17]Karayiannis, N.C., Giannousaki, A.E., Mavrantzas, V.G., and Theodorou, D.N. "Atomistic
Monte Carlo Simulation of Strictly Mono-disperse Long Polyethylene Melts through a
Generalized Chain Bridging Algorithm". J. Chemical Physics Vol: 117pp 5465-5472, 2002.
[18]Konrad, O., Lankau, T. "Solubility of Methane in Water: The Significance of the Methane-
Water Interaction Potential" J. Phys. Chem. B, Vol: 109 pp 23596-23604, 2005.
49

[19]Lacks, D.J. "First-order Amorphous-amorphous Transformation in Silica" Physical Review
E, Vol: 84:20 pp 4629-4632, 2000.
[20]Loucks, R.G., Reed, R.M., Ruppel, S.C., and Jarvie, D.M. "Morphology, Genesis, and
Distribution of Nanometer-Scale Pores in Siliceous Mudstones of the Mississippian
Barnett Shale" J. of Sedimentary Research, Vol: 79 pp 848-861, 2009.
[21]Luffel, D.L. and Guidry, F.K. "New Core Analysis Methods for Measuring Rock Properties
of Devonian Shale" J. Petroleum Tech. pp 1184-1190, 1992.
[22]Luffel, D.L., Hopkins, C.W., and Schettler, P.D. "Matrix Permeability Measurement of Gas
Productive Shales" Paper SPE-26633 presented at the SPE 68th Annual Tech. Conference
& Exhibition, Houston, Texas, October 3-6, 1993.
[23]Mancini, E. A., Goddard D. A., Barnaby R., and Aharon P. "Basin Analysis and Petroleum
System Characterization and Modeling, Interior Salt Basins, Central and Eastern Gulf of
Mexico" U.S. Department of Energy Final Technical Report, Phase 1, Project Number DE-
FC26-03NT15395: 422, 2006.
[24]Mavor, Matthew J., Hartman, Chad and Pratt, Timothy J. "Uncertainty in Sorption
Isotherm Measurements" paper 411, Proceedings 2004 International Coalbed Methane
Symposium, University of Alabama. Tuscaloosa May 2004.
[25]Marcus G. Martin, "MCCCS Towhee (Monte Carlo)" at:
http://towhee.sourceforge.net/algorithm/montecarlo.html, 2008.
[26]McCain, W. D. "The Properties of Petroleum Fluids", PennWell Books, 2nd Edition, Tulsa,
1990.
[27]Nath, S.K., Escobedo, F.A. and de Pablo J.J. "On the Simulation of Vapor-liquid Equilibria
for Alkanes" J. Chemical Physics Vol: 108:23 pp 9905-9911, 1998.
50

[28]Pollastro, R.M. "Total Petroleum System Assessment of Undiscovered Resources in the
Giant Barnett Shale Continuous (Unconventional) Gas Accumulation, Fort Worth Basin,
Texas", AAPG Bulletin 91 (4): pp 551-578, 2007.
[29]Roselman, I.C., and Tabor, D. "The friction of Carbon fibres" Journal of Physics D: Applied
Physics Vol: 9 pp 2517, 1976.
[30]Reynolds, M.M. and Munn, D.L. "Development Update for an Emerging Shale Gas Giant
Field Horn River Basin, British Columbia, Canada". Paper SPE 130103 presented at the
SPE Unconventional Gas Conference, Pittsburg, PA, 23-25 February 2010
[31]Schowalter, T. "Mechanics of secondary hydrocarbon migration and entrapment"
AAPG Bulletin, 63:5, pp 723-760, 1979.
[32]Sondergeld, C.H., Ambrose, R.J., Rai, C.S. and Moncrieff, J. "Micro-Structural Studies of
Gas Shales, Paper SPE 131771 presented at the SPE Unconventional Gas Conference,
Pittsburg, PA, 23-25 February 2010.
[33]Stapleton, M. R.; Panagiotopoulos, A. Z. "Application of excluded volume map sampling
to phase equilibrium calculations in the Gibbs ensemble" The Journal of Chemical
Physics, Volume 92, Issue 2, January 15, 1990, pp.1285-1293
[34]Smith, W., Forester, T. R., and Todorov, I. T., "The DL POLY 2 User Manual". STFC
Daresbury Laboratory, Cheshire, UK, 2008.
[35]Wang, F. P. and Reed, R. M. "Pore Networks and Fluid Flow in Gas Shales". Paper SPE
124253 presented at the Annual Technical Conference and Exhibition, SPE, New Orleans,
LA, October 4-7, 2009.
[36]Werder, T., Walther, J.H., Jaffe, R.L., Halicioglu, T., and Koumoutsakos, P. "On Water-
carbon Interaction for Use in Molecular Dynamics Simulations of Graphite and Carbon
Nanotubes", J. Phys. Chem. B, Vol: 107 pp 1345-1352, 2003.
51

[37]Widom, B. "Some Topics in the Theory of Fluids" J. Chemical Physics, Vol: 39:11 pp 2808-
2812, 1963.
[38]Zwanzig, R.W. "High-Temperature Equation of State by a Perturbation Method 1. Non-
polar Gases" J. Chemical Physics, Vol: 22:8 pp 1420-1426, 1954.

52

NOMENCLATURE
A
s
surface area, nm
2
B
g
gas formation volume factor, reservoir volume/surface volume
B
o
oil formation volume factor, reservoir volume/surface volume
B
w
water formation volume factor, reservoir volume/surface volume
c graphite pore wall wettability parameter for the graphite
G
st
total gas storage capacity, scf/ton
G
f
free gas storage capacity, scf/ton
G
a
adsorbed gas storage capacity, scf/ton
G
sL
Langmuir storage capacity, scf/ton
G
so
dissolved gas-in-oil storage capacity, scf/ton
G
sw
dissolved gas-in-water storage capacity, scf/ton
H pore size,
k
B
Boltzmann constant, 1.380650310
-23
kJ/K
-1

k
H
Henrys law constant
L Ostwald coefficient
L
x
length of computational cell in x-direction,
L
y
length of computational cell in y-direction,
apparent natural gas molecular weight, lbm/lbmole
M
CH4
molecular weight of methane, g/gmole
`
M
53

n number of moles
n
1
Gibbs isotherm number of moles at the start of pressure step, lb-moles
n
2
Gibbs isotherm number of moles at the end of pressure step, lb-moles
p pressure, psia
p
L
Langmuir pressure, psia
p
r1
pressure of reference cell at start, psia
p
r2
pressure of reference cell at end, psia
p
s1
pressure of sample cell at start, psia
p
s2
pressure of sample cell at end, psia
R universal gas constant, psia-ft
3
/mole-K
R
so
solution gas-oil ratio, scf/STB
R
sw
solution gas-water ratio, scf/STB
S
g
gas saturation, dimensionless
S
ge
effective gas saturation, dimensionless
S
o
oil saturation, dimensionless
S
w
water saturation, dimensionless
T reservoir temperature,
o
F
T
r1
reference cell temperature at start,
o
R
T
r2
reference cell temperature at end,
o
R
T
s1
sample cell temperature at start,
o
R
54

T
s2
sample cell temperature at end,
o
R
U Coulombic Potential, kJ/mole
V
b
bulk volume, ft
3

V
g
grain volume, ft
3

V
r
reference volume, ft
3

V
v
void volume, ft
3

V
ve
effective void volume, ft
3

V
v0
initial void volume, ft
3

V
v1
void volume step 1, ft
3

V
v2
void volume step 2, ft
3

z compressibility factor
z
r
compressibility factor of reference cell
Az interval across the pore space used for number density calculations,
Z
T,N
configurational partition function at temperature T and volume V
GREEK SYMBOLS
depth of the potential well, kJ
total porosity fraction, dimensionless

a
sorbed phase porosity fraction, dimensionless

e
effective porosity fraction, dimensionless

b
bulk rock density, g/cm
3

55

CH4
mass density of methane in pore, g/cm
3

f
free gas phase density, g/cm
3

Number
number density of methane, number of molecules/
3

s
sorbed phase density, g/cm
3
distance at which the inter-molecular potential is zero, nm
l *
2


excess chemical potential of the solute

Вам также может понравиться