Sei sulla pagina 1di 173

(/

A STUDY OF IMPROVED RECOVERY BY VAPORIZATION/

CONDENSATION PROCESS DUE TO ELEVATING

TEMPERATURES IN HYDROCARBON RESERVOIRS

by

VISHOK K. JAIN, B.E.

A THESIS

IN

PETROLEUM ENGINEERING

Submitted to the Graduate Faculty


of Texas Tech University in
Partial Fulfillment of
the Requirements for
the Degree of
MASTER OF SCIENCE

IN

PETROLEUM ENGINEERING

^ Approved

May, 1998
i^^^o ACKNOWLEDGEMENTS

I wish to express sincere thanks to Dr. Scott M. Frailey, Assistant Professor


at the Department of Petroleum Engineering, for his guidance throughout the
duration of the research, as well as for helping me in formulating a research
topic. Equally important to me is all the moral support and timely help he has
provided me during my graduate work at Texas Tech University.
I would like to express my deep gratitude to Dr. John J. Day, Professor
and Chairperson of the Department of Petroleum Engineering, for the financial
assistance in the form of a Teaching/Research Assistantship which made this
research work possible.
A very special thanks is extended to the Computer Modelling Group
(CMG), Calgary, Canada, for providing their phase-property program WinProp
free of cost for this study. In particular, I would like to thank Mr. Jim Erdle and
Mr. Mohamed Hassam for helping me in understanding the software.
I would like to thank Dr. Lloyd R. Heinze and Dr. M. Arnold for their
encouragement and for making valuable suggestions. I thank Johnita, Donna,
and Melissa who provided me with friendship and moral support for all these
years, and helped me out with the paperwork in the university.
I would like to thank all my friends in the department, who mc.de my stay
at Texas Tech so cheerful and fruitful. I would like to thank all my other friends
for their friendship, support and all the help they provided.
My sincere and deepest gratitude goes to my parents whose love and
caring support throughout my life encourage me to pursue my dreams and to
whom this work is dedicated.

11
TABLE OF CONTENTS

ACKNOWLEDGEMENTS ii
ABSTRACT v
LIST OF TABLES vii
LIST OF FIGURES viii
NOMENCLATURE xi
CHAPTER
1. INTRODUCTION 1
2. LITERATURE REVIEW 5
2.1 Basic Concepts of Phase Behavior 5
2.1.1 One-Component System 6
2.1.2 Two-Component System 8
2.1.3 Effect of Composition 9
2.1.4 Multi-Component System 10
2.2 Vapor-Liquid Phase Equilibria 12
2.2.1 Equilibrium Ratios 12
2.2.2 Flash Calculations 14
2.2.3 Separator Processes 16
2.3 Equation of State 18
2.3.1 Cubic Equations of State 20
2.3.2 Redlich-Kwong EOS and its Modifications 22
2.3.3 Peng-Robinson EOS and its Modifications 28
2.3.4 Schmidt and Wenzel EOS 31
2.3.5 Patel and Teja EOS 34
2.4 Applications of Equations of State 37
2.4.1 Determination of Equilibrium Constants 38

111
2.4.2 Simulation of Constant Volume Depletion Test 39
2.4.3 Simulation of Constant Composition Expansion Test 42
2.4.4 Simulation of Differential Liberation Test 44
2.5 Comparison of Equation of State 46
2.6 Tuning of Equation of State 47
2.7 Characterization of Plus Fractions 49
3. METROLOGY 59
3.1 Setting-Up the Initial Data and Characterization of Plus
Fractions 60
3.2 Selection of Equation of State Model 61
3.3 Characterization of Vaporization/Condensation Process 63
3.4 Assumptions 66
4. DISCUSSION OF RESULTS 73
4.1 Plus Fraction Characterization Results 73
4.2 Selection of Equation of State model 74
4.3 Results from Characterization of Vaporization/condensation
Process 75
5. CONCLUSIONS 149
6. RECOMMENDATIONS 150
REFERENCES 151

IV
ABSTRACT

A simulation of vaporization/condensation process was conducted to


study the effect of the process on the gas enrichment and overall recovery of a
low density crude oil.
The simulation begins with the identification of the most suitable equation
of state model to characterize vaporization/condensation process. This was
accomplished by simulating the constant volume depletion test and constant
composition expansion test using three equation of state models and comparing
the simulation results with the available experimental data. The equation of
model, which predicted the constant volume depletion data most accurately, was
selected to be used further in the study. The Peng-Robinson equation of state
was found to be most accurate in predicting the experimental data.
The vaporization/condensation process was characterized by employing
a combination of constant volume depletion test simulation and flash
calculations using the selected equation of state. The entire characterization of
vaporization/condensation process was performed on WinProp, a phase
property program.
The vaporization/condensation process was found to be feasible as it was
resulting in gas enrichment and increase in liquid saturation. These results were
much more pronounced when conducted the vaporization/condensation at the
high pressure and high vaporization temperature. Therefore, it was concluded
that the vaporization/condensation may result in additional recovery if carried
out at high reservoir pressure and high vaporization temperature.
This study was conducted on two samples of retrograde condensate fluid
and only three equation of state model were tested. Future work needs to be
done on other types of reservoir fluids and other more recent equations of state
such as Patel-Teja and Schmidt-Wenzel EOS should be verified.
The effects of flow characteristics of the fluid and heating mechanism
(steam injection, in-situ combustion) in the reservoir should be incorporated in
the future studies.

VI
LIST OF TABLES

3.1 Original Composition of Samples One and Two 68


3.2 Properties of Plus Fractions 68
3.3 Experimental Data for Sample 1 69
3.4 Experimental Data for Sample 2 70
4.1 Composition and Properties of SCN Fractions (Sample 1) 81
4.2 Composition and Properties of SCN Fractions (Sample 2) 81
4.3 Comparison of Retrograde Liquid Volume Calculated By Different
EOS (Sample 1) 82
4.4 Comparison of Retrograde Liquid Volume Calculated By Different
EOS (Sample 2) 83
4.5 Comparison of Produced Gas Calculated By Different EOS (Sample 1) 84
4.6 Comparison of Produced Gas By Different EOS (Sample 2) 85
4.7 Comparison of Saturation Pressures Calculated By Different EOS 85
4.8 Comparison Between Experimentally Obtained Composition and
Composition Calculated By PREOS at Depleted Reservoir Pressure
(Sample 1) 86
4.9 Comparison Between Experimentally Obtained Composition and
Composition Calculated By PREOS at Depleted Reservoir Pressure
(Sample 2) 87
4.10 Calculated Compositions at Depleted Reservoir Pressures and
Original Reservoir Temperature (sample 1) 88
4.11 Calculated Compositions at Depleted Reservoir Pressures and
Original Reservoir Temperature (sample 1) 89
4.12 Moles of Fluid in Place After each Cycle 90

Vll
LIST OF FIGURES

2.1 P-V-T Diagram for a Single-Component System 51


2.2 Pressure-Temperature Diagram for a Single-Component System 52
2.3 Pressure-Volume Diagram for a Single-Component System 53
2.4 Typical Pressure-Temperature Diagram for a Binary System 54
2.5 Effect of Composition on Phase Envelope for a Binary System 55
2.6 Location of Critical Loci for Several Binary Systems 56
2.7 Typical Pressure-Temperature Diagram for a Multi-Component
System 57
2.8 The Exponential and Left-Skewed Distribution Functions 58
3.1 Flow Diagram of Procedure Used in the Study 71
3.2 Schematic Diagram of Vaporization/Condensation Process 72
4.1 Comparison of Relative Volume Data (Sample 1, PREOS) 91
4.2 Comparison of Relative Volume Data (Sample 1, SRKEOS) 92
4.3 Comparison of Relative Volume Data (Sample 1, SRKGDEOS) 93
4.4 Comparison of Relative Volume Data (Sample 2, PREOS) 94
4.5 Comparison of Relative Volume Data (Sample 2, SRKEOS) 95
4.6 Comparison of Relative Volume Data (Sample 2, SRKGDEOS) 96
4.7 Comparison of Z-Factor Data (Sample 1, PREOS) 97
4.8 Comparison of Z-Factor Data (Sample 1, SRKEOS) 98
4.9 Comparison of Z-Factor Data (Sample 1, SRKGDEOS) 99
4.10 Comparison of Z-Factor Data (Sample 2, PREOS) 100
4.11 Comparison of Z-Factor Data (Sample 2, SRKEOS) 101
4.12 Comparison of Z-Factor Data (Sample 2, SRKGDEOS) 102
4.13 Comparison of CVD Test Data (Sample 1, PREOS) 103
4.14 Comparison of CVD Test Data (Sample 1, SRKEOS) 104

Vlll
4.15 Comparison of CVD Test Data (Sample 1, SRKGDEOS) 105
4.16 Comparison of CVD Test Data (Sample 2, PREOS) 106
4.17 Comparison of CVD Test Data (Sample 2, SRKEOS) 107
4.18 Comparison of CVD Test Data (Sample 2, SRKGDEOS) 108
4.19 Variation in Mole % of Liquid and Vapor (Sample 1, 750 psia, 700 F 109
4.20 Variation in Mole % of Liquid and Vapor (Sample 1, 750 psia, 500 F 110
4.21 Variation in Mole % of Liquid and Vapor (Sample 1, 250 psia, 700 F 111
4.22 Variation in Mole % of Liquid and Vapor (Sample 1, 250 psia, 500 F 112
4.23 Variation in Mole % of Liquid and Vapor (Sample 2, 750 psia, 500 F 113
4.24 Variation in Mole % of Liquid and Vapor (Sample 2, 750 psia, 350 F 114
4.25 Variation in Mole % of Liquid and Vapor (Sample 2, 250 psia, 500 F 115
4.26 Variation in Mole % of Liquid and Vapor (Sample 2, 250 psia, 350 F 116
4.27 Variation in Volume % of Liquid and Vapor
(Sample 1, 750 psia, 700 F) 117
4.28 Variation in Volume % of Liquid and Vapor
(Sample 1, 750psia, 500 F) 118
4.29 Variation in Volume % of Liquid and Vapor
(Sample 1, 250psia, 700 F) 119
4.30 Variation in Volume % of Liquid and Vapor
(Sample 1, 250psia, 500 F) 120
4.31 Variation in Volume % of Liquid and Vapor
(Sample 2, 750psia, 500 F) 121
4.32 Variation in Volume % of Liquid and Vapor
(Sample 2, 750psia, 350 F) 122
4.33 Variation in Volume % of Liquid and Vapor
(Sample 2, 250psia, 500 F) 123
4.34 Variation in Volume % of Liquid and Vapor
(Sample2, 250psia, 350 F) 124
4.35 Variation in Overall Composition (Sample 1, 750 psia, 700 F) 125
4.36 Variation in Overall Composition (Sample 1, 750 psia, 500 F) 126

IX
4.37 Variation in Overall Composition (Sample 1, 250 psia, 700 F) 127
4.38 Variation in Overall Composition (Sample 1, 250 psia, 500 F) 128
4.39 Variation in Overall Composition (Sample 2, 750 psia, 500 F) 129
4.40 Variation in Overall Composition (Sample 2, 750 psia, 350 F) 130
4.41 Variation in Overall Composition (Sample 2, 250 psia, 500 F) 131
4.42 Variation in Overall Composition (Sample 2, 250 psia, 350 F) 132
4.43 Variation in Liquid Composition (Sample 1, 750 psia, 700 F) 133
4.44 Variation in Liquid Composition (Sample 1, 750 psia, 500 F) 134
4.45 Variation in Liquid Composition (Sample 1, 250 psia, 700 F) 135
4.46 Variation in Liquid Composition (Sample 1, 250 psia, 500 136
4.47 Variation in Liquid Composition (Sample 2, 750 psia, 500 'F 137
4.48 Variation in Liquid Composition (Sample 2, 750 psia, 350 138

4.49 Variation in Liquid Composition (Sample 2, 250 psia, 500 F 139

4.50 Variation in Liquid Composition (Sample 2, 250 psia, 350 F 140

4.51 Variation in Vapor Composition (Sample 1, 750 psia, 700 F) 141

4.52 Variation in Vapor Composition (Sample 1, 750 psia, 500 F] 142

4.53 Variation in Vapor Composition (Sample 1, 250 psia, 700 F) 143

4.54 Variation in Vapor Composition (Sample 1, 250 psia, 500 F] 144

4.55 Variation in Vapor Composition (Sample 2, 750 psia, 500 F) 145

4.56 Variation in Vapor Composition (Sample 2, 750 psia, 350 F) 146

4.57 Variation in Vapor Composition (Sample 2, 250 psia, 500 F) 147

4.58 Variation in Vapor Composition (Sample 2, 250 psia, 350 F] 148


NOMENCLATURE

a Equation of state parameter


ai Parameter 'a' for component 'i'
aj Parameter 'a' for component 'j'
am Parameter 'a' for a mixture
ai-as Constants defined in equations 2.59 and 2.60
A Constant defined in equation 2.16
b Equation of state parameter
bi Parameter 'b' for component 'i
bm Parameter 'b' for a mixture
B Constant defined in equation 2.17
Bod Relative oil volume
c Parameter for Patel and Teja equation of state
Ci Volume correction parameter
Cm Parameter 'c' for a mixture
C Constant defined in equation 2.94
Ci-Ce Alkane series hydrocarbons with 1-6 carbon atoms
C7+ Hydrocarbons with 7 or more carbon atoms
CO2 Carbon dioxide
CVD Constant volume depletion
CCE Constant composition expansion
d Parameter defined in equation 2.98
EOS Equation of State
EXP(x) Equivalent to e"
f Fugacity
f(nv) Function defined in equation 2.4

XI
f(nv) Function defined in equation 2.6
Gp Gas produced, scf
(Gp)t Total cumulative gas produced, scf
Ki Equilibrium constant for component 'i'
KiA Assumed equilibrium constant of component 'i'
Kij Binary interaction coefficient between components 'i' and ']
Kij Parameter defined in equation 2.44
In Natural logarithm
m Parameter defined in equation 2.29
rrio Constant defined in equation 2.56
mi Constant defined in equation 2.71
m2 Constant defined in equation 2.72
MWc7+ Molecular weight of C7+ fractions
n Number of components in the mixture
ni Initial moles in place
nL Number of moles in the liquid phase
nv Number of moles in the vapor phase
np Number of moles produced
(nL)actual Actual number of moles in the liquid phase
(nL)in-place Moles of liquid in place
(nv)actual Actual number of moles in the vapor phase
^nv}produceci Moles of produced gas
(nv)n New value of nv
(nv)r Remaining moles of gas
N2 Nitrogen

P System pressure, psia


Pb Bubble point pressure, psia

Xll
pc Critical pressure, psia
pen Critical pressure of the component 'i', psia
pd Dew point pressure, psia
psat Saturation pressure, psia
PR Peng-Robinson
PT Patel-Teja
Q Parameter defined in equation 2.97
R Universal gas constant, 10.73 psia-ftV (lb-mole)(R)
Rsd Solution gas-oil ratio, scf/STB
Si Dimensionless shift parameter
SRK Soave-Redlich-Kwong
SRKGDEOS Soave-Redlich-Kwong equation of state with the constant 'a'
proposed by Graboski and Daubert
SW Schmidt-Wenzel
T System temperature, R
Tb Boiling point of the plus fraction, R
Tc Critical temperature, R
Tci Critical temperature of the component 'i', R

^ " Reduced temperature


^ Coefficient for SW EOS defined in equation 2.74
^ Volume, ft^/mole
^i Initial volume, ft^/mole
^^1 Relative phase volume
^sat Volume at saturation pressure, ft^/mole
^L Volume of the liquid phase
^g Volume of the vapor phase
^t Total volume of the fluid, ft^

XIU
W Coefficient for SW EOS defined in equation 2.75
Xi Mole fraction of the c o m p o n e n t ' i ' in the liquid phase
(xi)in-piace Molc fraction of the c o m p o n e n t 'i' in the liquid in place
yi Mole fraction of the c o m p o n e n t 'i' in the v a p o r p h a s e
(yi)produced Molc fraction of the c o m p o n e n t 'i' in the p r o d u c e d gas
Zi Mole fraction of c o m p o n e n t 'i' in the h y d r o c a r b o n mixture
Zj Mole fraction of c o m p o n e n t 'j' in the h y d r o c a r b o n mixture
Z Compressibility factor
Z^ Compressibility factor of the liquid phase
Z^ Compressibility factor of the vapor phase
ZRA Rackett compressibility factor
Zc Critical compressibility factor
a T e m p e r a t u r e d e p e n d e n t coefficient in SRK a n d PR EOS
tti Coefficient a for c o m p o n e n t ' i '
ttj Coefficient a for component']
pc Smallest positive root of equation 2.64
Y Specific gravity
yc7+ Specific gravity of h e p t a n e - p l u s fraction
e Error tolerance
i P a r a m e t e r defined in equation 2.45
^c Parameter defined in equation 2.65
(J). Fugacity coefficient of c o m p o n e n t ' i ' in a mixture
<j,.L Fugacity coefficient of t h e c o m p o n e n t 'i' in the liquid p h a s e
(j).v Fugacity coefficient of the c o m p o n e n t ' i ' in the v a p o r p h a s e
yi, Coefficient defined in equation 2.39
^. Coefficient defined in equation 2.38

CO
Acentric factor

XIV
coi Acentric factor for component 'i'
com Acentric factor for a mixture
Qa Constant in equation of state
Qb Constant in equation of state
Qc Constant in PT equation of state

XV
CHAPTER 1
INTRODUCTION

The predictions of the phase behavior and fluid properties of the


hydrocarbon systems are important in several different aspects of the petroleum
industry. These include the development of reservoirs containing black oil,
volatile oil, gas condensate, and natural gas, as well as the design of gas injection
processes, enhanced oil recovery schemes, and hydrocarbon processing units.
Traditionally, chemical and petroleum engineers engaged in the production of
oil and natural gases from reservoirs have not devoted much interest to details
regarding the properties and compositions of the hydrocarbon fluids produced.
Design of production schemes and equipment was mainly based on experience.
This was, however, insufficient considering the enormous investment required
for oil and gas exploration and production. Today, production systems and
equipment undergo detailed engineering analysis and careful design to optimize
the amount of hydrocarbons produced and minimize the operating costs and
investments.
Analysis and design of oil and gas production facilities and schemes
require thorough knowledge of the thermodynamic and physical properties of
the hydrocarbon fluids. To determine the volume of oil and gas a separator
produces under given conditions, it is necessary to know the vapor-liquid
relationships and the densities of the mixture. Understanding the flow process
in the reservoir requires the knowledge of viscosity and capillary pressure. Heat
exchanger design requires the knowledge of the thermal conductivity of these
fluids.
It would be best if knowledge of the above mentioned properties were
available from experimental observations. Accurate laboratory studies of PVT
properties and phase-equilibria behavior of reservoir fluids are, therefore,
necessary for characterizing these fluids and evaluating their volumetric
performance at various pressure and temperature conditions. There are many
laboratory analyses that can be made on a reservoir fluid sample. The amount of
data desired determines the number of tests performed in the laboratory. In
general, there are three types of laboratory tests used to measure hydrocarbon
reservoir samples.
1. Primary Tests: These routine tests involve the measurement of the
specific gravity and the gas-oil ratio of the produced hydrocarbon
fluids.
2. Detailed Laboratory Tests: These typical tests are performed to
characterize the hydrocarbon fluids. They involve the measurement of
the fluid composition, molecular weight, viscosity, compressibility,
saturation pressure, formation volume factor, solubility, differential
liberation, constant volume depletion, and constant composition
expansion characteristics.
3. Swelling Tests: In addition to primary and detailed laboratory tests,
which are fairly standard, another type of test may be performed for
very specific applications. If a reservoir is planned to be depleted
under a gas injection or dry gas cycling scheme, a swelling test should
be performed.
If reservoir fluid samples are available, the fluid properties of interest can
be measured with the above mentioned laboratory analysis procedures.
However, these analyses are usually conducted at reservoir temperatures only
and the variation of the properties with temperature is not available for
production system calculations. Also, in many cases a PVT analysis may not be
available early in the life of the reservoir or may never be available because of the
relatively high cost of performing these experimental PVT tests and the
uncertainties in the accuracy of such laboratory measurements. To overcome
these obstacles, empirical correlation methods have been developed for
predicting various fluid properties and describing phase behavior characteristics
of reservoir fluids from limited data. In recent years, these empirical correlation
methods found wide acceptability among petroleum engineers for numerical
simulation of experimental PVT tests. One such very popular method is to
employ equations of state to simulate laboratory tests.
An equation of state (EOS) is an analytical expression relating the pressure
(p) to the temperature (T) and the volume (V). Numerous equations of state have
been proposed and applied for phase behavior modelling and compositional
reservoir simulators. The most commonly used equations have been compared
by a number of investigators,^-^ resulting in a general conclusion that none of the
equations can be singled out as the most superior one to predict all the properties
at all conditions. It is, however, widely accepted that equations of state, when
properly tuned, are capable of adequately simulating the PVT properties of
reservoir fluids and can consequently save significant time and expense by
eliminating the need to perform a complete set of experimental PVT tests on each
and every new reservoir fluid. In addition, equations of state can also be used to
check the quality of the fluid sample collected and relative accuracy of many
laboratory tests.
The primary objective of this study was to prove or disprove the concept
of a vaporization/condensation process of increasing the temperature of a low
pressure hydrocarbon reservoir that contains low oil saturation. It was
conceptualized^o that when the immobile residual oil is subjected to increase in
temperature, a part of this oil converts into vapor and moves ahead in the
reservoir because of the fact that vapors have a much higher mobility compared
to liquids. These vapors moves ahead to the position of the reservoir which is
not affected by the increase in the temperature and become partly or fully
condensed due to a decrease in temperature (original reservoir temperature). As
the heated zone moves toward this area of newly condensed fluids, this zone is
again subjected to increase in temperature. This vapor again moves ahead to the
zone which is at the reservoir temperature and becomes condensed partially or
completely due to reduced temperature. This process was repeated until either
the oil was completely vaporized or the increase in the vapor content after each
stage (one vaporization/condensation cycle) was insignificant.
The scope of this study begins with the tuning of an equation of state to
match available experimental data. The tuned equation of state was then used to
model various increases and decreases in temperature of the immobile reservoir
fluid to predict the volumetric and phase behavior properties of the fluid.
Additionally, to achieve and further complement the primary objective,
the following secondary objectives were set:
to identify the most appropriate equation of state model that simulates
the available experimental PVT data.
to show that the vaporization/condensation process, if proved, results
in additional recovery by either enrichment of the produced gases or
mobilization of the oil due to increased saturation.
to observe the variation in the mole percent and volume percent of the
vapor and liquid phases with every cycle.
to study the variation in the composition of the liquid and vapor
phases with every cycle.
CHAPTER 2
LITERATURE REVIEW

Crude oil and natural gas are a complex mixture consisting


predominantly of naturally occurring hydrocarbons that may exist in the solid,
liquid, or gaseous states, depending upon the conditions of pressure and
temperature to which it is subjected. These hydrocarbons exist at elevated
temperature and pressure in the reservoir, and upon production of these
hydrocarbon mixtures at the surface, the temperature and pressure of the
mixture are reduced. The phase state of the mixture at the surface conditions
depends on the composition of the fluid as produced from the well and on the
temperature and pressure at which it is captured. The fluid remaining in the
reservoir at any stage of depletion also undergoes physical changes as the
reservoir pressure is reduced by production of oil and gas from the reservoir.
Under natural production, the reservoir temperature is assumed constant
through the life of the reservoir. It is necessary to study the physical properties
of these naturally existing hydrocarbons, and in particular, their variation with
changes in pressure and temperature of the system, to estimate the performance
of the reservoir.
The intent of this chapter is to review the concepts of phase behavior and
vapor-liquid phase equilibria. Additionally, various widely used equations of
state were reviewed and compared for their applicability and accuracy.

2.1 Basic Concepts of Hvdrocarbon Phase Behavior^^' ^2, i3,i4


A ''phase" is defined as that part of the system that is uniform in physical
and chemical properties, homogeneous in composition, and separated from other
coexisting phases by definite boundary surfaces. For example, ice, water, and
steam constitute three separate phases of the pure substance H2O because each is
homogeneous and physically distinct from the others and each is clearly defined
by the boundaries existing between them.
There are three different familiar phases: solid, liquid, and gas (vapor). A
solid possesses a definite shape and is hard to the touch. It is composed of
molecules with very low energy that stay in one place even though they vibrate.
A liquid has a definite volume but no definite shape. It will assume the shape of
the container in which it is placed but will not necessarily fill the container. The
molecules of which the liquid is composed possess more energy than a solid. A
vapor has no definite volume or shape and will fill a container in which it is
placed. The molecules have more energy than in the liquid form.
Whether a substance exists in a solid, liquid, or vapor phase is determined
by the temperature and pressure acting on the substance. It is known that solid
phase (ice) can be changed to liquid phase (water) by increasing its temperature
and this liquid phase can be changed to vapor phase (steam) by increasing the
temperature further. This change in phases is termed as Phase Behavior.
The objective of this section is to review the basic principles of
hydrocarbon phase behavior and describe the use of phase diagrams in
characterizing the volumetric behavior of pure substances (single-component
system) and mixtures (multi-component system). A phase diagram is a record of
the effects of pressure, temperature, and the composition on the type and
number of phases that can coexist in equilibrium with each other.

2.1.1 One-Component Svstem


The simplest type of hydrocarbon mixture is one that contains only one
component. A single component system is composed entirely of one kind of
atom or molecule. The word ''pure" is generally used to describe a single
component system. The qualitative understanding of the relationship between
temperature (T), pressure (p), and volume (V) of pure components can provide
an excellent basis for understanding the phase behavior of complex petroleum
mixtures. Figure 2.1 is a typical phase diagram for a pure substance. It has three
axes p, V, and T and is composed of a series of plane surfaces, each of which
represents a given phase or mixture of phases. For example, plane BDHG
represents solid plus liquid phase, plane FGIJ represents solid plus vapor, and
plane HCI for the liquid plus vapor. All these planes are perpendicular to the
temperature axis.
A three-dimensional phase diagram like Figure 2.1 is difficult to use and
generally not required to describe phase behavior in hydrocarbon systems where
the vapor-liquid region (plane HCI) of the phase diagram is of primary concern.
Therefore, a projection of the three dimensional diagram, like P-T and P-V
diagram, is sufficient to describe the phase behavior of the hydrocarbon systems.
Figure 2.2 is a Typical P-T diagram for single component system. Since all of the
two-phase planes in figure 2.1 are perpendicular to the T axis, they appear as
single line in a P-T diagram. The resulting curve (line HC), which terminates at
the critical point (point C), is the dividing line between the area where liquid and
vapor exists. The curve is commonly called the "vapor pressure curve" or the
"boiling point curve." The corresponding pressure at any point on the curve is
called the vapor pressure.
Figure 2.3 is a typical pressure-volume diagram for a single component
system and is obtained experimentally by placing a fixed quantity of pure
component in a cylinder fitted with a frictionless piston and subjecting the
component to changes in pressure and temperature. The initial condition is
represented by point E on this diagram. The pure component which was in the
vapor state in the initial conditions (point E, temperature Ti) was subjected to a
isothermal pressure increase by forcing the piston into the cylinder. This results
in a decrease in the gas volume until it reaches the point F in the diagram, where
the liquid begins to condense. The corresponding pressure is known as the dew
point pressure pd, and is defined as the pressure at which the first drop of liquid
is formed. Further increase in pressure results in additional condensation. This
condensation process is characterized by a constant pressure and represented by
horizontal line FG. At point G, only traces of gas remain and the corresponding
pressure is called the bubble point pressure pb, and is defined as the pressure at
which the first sign of gas formation is detected. The slight forcing of the piston
into the cylinder at point G results in a sharp increase in pressure (point H)
without an appreciable decrease in the liquid volume. This is because the
compressibility of the liquid phase is much less compared to the vapor phase.
By repeating the above procedure at progressively increasing
temperatures, a family of curves of equal temperatures (isotherms) is
constructed. The dashed curve connecting the dew points is called the dew point
curve (line FC) and the dashed curve connecting the bubble points is called the
bubble point curve (line GC). A characteristic of a single component system is
that at a given temperature, the dew point and the bubble point pressure are
equal. These two curves meet at point C which is known as critical point. The
corresponding pressure and volume are called the critical pressure pc and critical
volume Vc respectively. For a single component system, the critical point is
defined as the highest value of pressure and temperature at which two phases
can coexist. The area enclosed by the phase envelope AFCGB is called the two
phase region. Within this region, vapor and liquid can coexist in equilibrium.
Outside the phase envelope, only one phase can exist.
2.1.2 Two-Component system
A distinguished feature of a single-component system is that, at a given
temperature, two phases (liquid and vapor) can exist in equilibrium at only one
pressure. But for a two-component or binary system, two phases can exist in
equilibrium at various pressures at the same temperature. The phase behavior
description of a binary system is more complex compared to a one-component
system because of the addition of another variable, composition, to the system.
The thermodynamic and physical properties of a binary system vary with a
change in composition. The effect of this variable is seen by comparing the
pressure-temperature diagram for a single component system (Figure 2.2) with
that of binary system (Figure 2.4). Figure 2.4 indicates that the pressure-
temperature relationship can no longer be represented by a simple vapor
pressure curve as in the case of single-component system but takes on the form
of a phase envelope ACB. This phase envelope is formed by an intersection of
the dew point curve (line BC) and the bubble point curve (line AC) at the critical
point C. The dashed line within the phase envelope are called "quality lines"
and they represent the pressure and temperature conditions of equal volumes of
liquid. For each possible composition, a distinct phase diagram exists. The
critical point for a binary system is defined as the point at which all intensive
properties of the vapor and liquid phases are equal. An intensive property is one
that has the same value for any part of a homogeneous system as it does for the
whole system, i.e., a property which is independent of the quantity of the system.
Critical pressure, critical temperature, density, composition, and viscosity are
examples of intensive properties.
2.1.3 Effect of Composition
Figure 2.5 demonstrates the effect of changing the composition of the
binary system on the shape and location of the phase envelope. Two lines
that form the phase envelope are the vapor pressure curves for methane and
propane. Three phase envelopes are shown for three different compositions of a
methane and propane mixture. These envelopes pass continuously from the
vapor pressure of the one pure component to that of the other as the composition
is varied. The dashed line is the line drawn tangent to all possible phase
envelopes at the critical point on each curve. It illustrates the locus of critical
points for the binary system. It can be seen from Figure 2.5 that when one of the
constituents becomes more dominant, the binary system tends to exhibit a
relatively narrow phase envelope and displays critical properties close to the
dominant component. The size of the phase envelope enlarges as the
composition of the mixture becomes evenly distributed between the two
components.
Figure 2.6 shows the critical loci for a number of common binary systems.
It can be seen that the critical pressure of the mixtures is considerably higher
than the critical pressure of the components in the mixtures. The greater the
difference in the boiling point of the two components, the higher the critical
pressure of the mixture will be.

2.1.4 Multi-Component Svstem


The phase behavior of a multi-component system in the liquid-vapor
region is very similar to that of a binary system. However, as the system
becomes more complex with a greater number of components, the pressure and
temperature ranges in which two phases can coexist increases significantly.

10
Figure 2.7 shows a typical pressure-temperature phase diagram of a multi-
component system with a specific overall composition. Although a different
hydrocarbon system would have a different phase diagram, the general
configuration is similar.
These multi-component pressure-temperature phase diagrams are
commonly used to:
classify reservoirs
classify the naturally occurring hydrocarbon systems
describe the phase behavior of the reservoir fluids.
To completely understand the significance of the p-T diagram (Figure 2.7),
it is necessary to identify and define following key points on the p-T diagram:
Cricondentherm Tct (point E) - The cricondentherm is defined as the
maximum temperature above which liquid can not be formed
regardless of pressure. The corresponding pressure is known as
cricondentherm pressure pct.
Cricondenbar pcb, (point D) - The cricondenbar is the maximum
pressure above which no gas can be formed regardless of temperature.
The corresponding temperature is called the cricondenbar temperature
Tcb.

Critical point (point C) - The critical point for a multi-component


system is referred to as the state of pressure and temperature at which
all intensive properties of the vapor and liquid phases are equal. At
the critical point, the corresponding pressure and temperature are
called the critical pressure pc and critical temperature Tc of the system.
Phase Envelope (line BCA) - The region enclosed by bubble point
curve (line BC) and dew point curve (line AC) is called the phase

11
envelope of the system. Two phases (liquid and vapor) can coexist in
equilibrium only inside the phase envelope.
Quality Lines - The dashed lines within the phase diagram are called
quality lines. They describe pressure and temperature conditions for
equal volume of liquids. All quality lines converge at the critical point.

2.2 Vapor-Liquid Phase Equilibria i^> ^^' ^^' ^^


A complex mixture of hydrocarbon compounds or components can
exist as a single phase liquid, a single-phase gas or as a two-phase mixture
depending on the pressure, temperature and composition of the mixture. The
conditions under which these different phases can exist is a matter of
considerable practical importance in designing surface separation facilities and
developing compositional reservoir simulator models. The quantitative analysis
of a two-phase system involves the determination of the mole fractions of vapor
and liquid present at a given condition and computation of the number of
coexisting phases. These type of calculations are based on the concept of
equilibrium ratios.

2.2.1 Equilibrium Ratios


The distribution of a component of a system between vapor and liquid is
expressed by the equilibrium ratio, also referred to as a equilibrium constant.
The equilibrium ratio, Ki, of a given component is defined as the ratio of the mole
fraction of the component in the vapor phase, yi, to the mole fraction of the
component in the liquid phase, Xi. Mathematically, the relationship is expressed
as:

K^ = (2.1)

12
The K-values of each component in a system is a function of the pressure,
temperature, and composition of the system. The composition of the system
considerably affects the K-values above 1000 psi; below 1000 psi the effect is
relatively small. The K-values can be physically measured for each reservoir
fluid, calculated from empirical correlations, or selected from a large number of
K-value measurements available in the literature. Numerous methods have been
proposed for predicting the equilibrium ratios of the hydrocarbon mixtures.
These correlations range from simple mathematical expressions to complicated
expressions containing several compositional dependent variables. Among these
methods, the most widely used method is Wilson's correlation.^^
Wilson^^ proposed the following expression for estimating K-values at
low pressures:

n 5.37(l+a>,.)fl-y
K^=^EXP^ ^ '^^ (2.2)
P
Where, pd and Td are the critical pressure and the critical temperature of
the component i. T and p are the system temperature and system pressure, and
coi is the acentric factor of the component i. The acentric factor is a correlating
parameter used to characterize the acentricity or non-sphericity of a molecule.
For detailed procedure of other methods of calculating equilibrium ratios,
the reader may consult the following textbooks.
Hydrocarbon Phase Behavior by Tarek Ahmed.^^
Gas Conditioning and Processing by John M. Campbell.^2
The main disadvantage of Wilson's correlation and other methods is their
limited range of pressure and temperature within which they can provide
accurate results, and their inability to incorporate the effect of the other
components present in the mixture on the prediction of K-values. These
disadvantages can be overcome by using equation of state methods for

13
predicting equilibrium ratios. This method provides accurate prediction of K-
values over a larger range of pressure and temperature for complex hydrocarbon
mixtures. The main disadvantage of this method is that it is very complex and
normally requires use of computers. A detailed description of using an equation
of state to predict equilibrium ratios is presented in next chapter.
The vast amount of work that has been done on equilibrium ratio studies
indicates their importance in solving phase behavior problems in reservoir and
process engineering. Equilibrium ratios are commonly used in the petroleum
industry to perform flash and separator calculations and to determine dew-point
and bubble-point pressures.

2.2.2 Flash Calculations


Flash calculations are an integral part of all reservoir and process
engineering calculations. They are required whenever it is desirable to know the
amount (in moles) of hydrocarbon liquid and vapor coexisting in a reservoir or a
vessel at a given pressure and temperature. These calculations are also
performed to determine the composition of the existing hydrocarbon phases.
Flash calculations are used to determine:
the number of phases,
molar volume of each phase,
molar composition of each phase.
A step by step procedure of performing flash calculations on the basis of
one mole of the hydrocarbon mixture is as follows:
Stepl: Calculation of moles of vapor phase, nv
Assume a starting value of nv. A good starting value can be calculated
from the following expression:

14
ny = (2.3)
A-B

where
i *
-.(^,-lJ
'Z,(A:,-I)"
i

where Zi is the mole fraction of the component 'i' in the hydrocarbon


mixture.
Evaluate the function f(nv) as given by the following expression using
the assumed value of nv.

/K) = Z ny{K,-\) +\
=0 (2.4)

If the absolute value of function f(nv) is smaller than a preset tolerance,


e.g., 10'^^ than the assumed value of nv is the desired solution.
If the absolute value of the function is greater than the preset tolerance,
then a new value of nv can be calculated from the following
expression:

( \ -^M (2.5)
(ny j =ny - (v-

where,
(nv)n = new value of nv

^,(^,-1)'
/(.)=-I (2.6)

the above procedure is repeated with the new values of nv, until
convergence is achieved.
Step 2: Calculation of moles in liquid phase nL from the following

15
expression:
,,=!-, (2.7)
Step 3: Calculation of composition of liquid phase Xi from the following
equation:

^L -^nyK,

Step 4: Calculation of composition of vapor phase yi from the following


equation:

^ - ^ . ^ ' ' ^

2.2.3 Separator Processes


Produced reservoir fluids are complex mixtures of components having
different physical properties. As a crude oil mixture flows from a high pressure,
high temperature petroleum reservoir to the surface, it experiences a reduction in
pressure and temperature. Because of this reduction in pressure and
temperature, gases evolve from the liquid and the well stream changes in
character. The physical separation of these phases, and the manner in which
they are separated influences the stock tank oil recovery. The principal means of
surface separation of oil and gas is the conventional stage separation. The stage
separation is the process in which gaseous and liquid hydrocarbon are separated
into vapor and liquid phases by two or more separators. These separators are
usually operated in series of consecutive lower pressures. Each condition of
pressure and temperature at which hydrocarbon phases are flashed is called a
stage of separation. There are two types of gas-oil separation process:
differential separation,
flash separation.

16
The differential separation process is generally performed for a black oil to
simulate the conditions encountered in the reservoir. The sample of reservoir
fluid in the laboratory cell is brought to the bubble point pressure, and the
temperature is set to the reservoir temperature. Pressure is reduced by
increasing the cell volume. All the gas is expelled from the cell while the
pressure is held constant by reducing the cell volume. The process is repeated in
steps until atmospheric pressure is reached. The solution gas that is liberated
from an oil sample during a decline in pressure is removed continuously from
contact with the oil before establishing equilibrium with the liquid phase. This
type of separation is characterized by the varying composition of the total
hydrocarbon system. The similar process used to simulate the reservoir
conditions for gas condensate is called as constant volume depletion. The only
difference is that in the constant volume depletion process, the sample of the
reservoir fluid is brought to the dew point pressure and the reservoir
temperature, and then the pressure is reduced in steps.
In the flash liberation process, the gas which is liberated from the oil
during the pressure decline remains in contact with the oil from which it was
liberated. The overall system composition remains constant during this type of
liberation because no gas is removed from the PVT cell during pressure
reductions. This procedure is sometimes referred in the literature as flash
vaporization, constant composition expansion, or flash expansion.
The differential liberation test is considered to better describe the
separation process taking place in the reservoir and is also considered to better
simulate the flowing behavior of hydrocarbon systems at conditions above the
critical gas saturation. As the saturation of the liberated gas reaches the critical
gas saturation, the liberated gas begins to flow, leaving behind the oil that
originally contained it. This is attributed to the fact that gases have, in general,

17
higher mobility than oils. The fluids produced from the reservoir to the surface
are considered to undergo flash liberation process.
These processes can also provide considerable insight in describing PVT
relationship for complex hydrocarbons. The experimental data obtained from a
flash liberation test include:
bubble point pressure
isothermal compressibility coefficient of the liquid phase above the
bubble point pressure
the two phase volume as a function of pressure, below the bubble
point.
The experimental data obtained from the differential liberation test
include:
amount of gas in solution as a function of pressure
the shrinkage in the oil volume as a function of pressure
properties of the evolved gas including the composition,
compressibility factor, and gas specific gravity
density of the remaining oil as a function of pressure.
The procedure for numerical simulation of differential and flash liberation
using equation of state is discussed in section 2.4.

2.3 Equation of State


An equation of state (EOS) for any substance is a mathematical expression
that provides the relationship between the pressure (p), volume (V), and the
temperature (T) of that substance. Equations of state are very sophisticated tools
for engineering applications and commonly used in the petroleum industry for
volumetric and phase behavior modelling of reservoir fluids.

18
The development of an EOS for the description of real systems started in
1893 with the publication of van der Waal's equation of state. Since then
hundreds of equations of state have been published in the literature and
promising new equations keep appearing. Leland^^ (1980) conducted a
comprehensive review of equations of state and classified them into the
following four families:
van der Waal family (vdW)
Benedict-Webb-Rubin family
reference-fluid equations
augmented-rigid-body equations.
The van der Waal family mainly encompasses simple, mostly cubic
equations of state. Their main characteristic is the separation between the
repulsive and attractive molecular forces. Leland reported that vdW EOS,
despite their simplicity, display quantitatively accurate performances in
describing the PVT relationship for complex hydrocarbon fluids.
The Benedict-Webb-Rubin (BWR) family includes complicated equations
of state that are empirical extension of the virial EOS. Besides the Benedict-
Webb-Rubin^^ EOS itself, the most significant members of this family for oil and
gas applications are the Sterling^^ EOS, and the Lee and Kesler^o EOS.
The reference-fluid equations are generally focused on accurate
representation of a large amount of PVT data for pure substances. An important
application of reference-fluid equations is their use as reference fluids in
corresponding state theories.
The augmented-rigid-body family combines the description of repulsive
forces between hard molecules of different shapes with the expression for the
molecular attractions. The sound theoretical basis of the repulsive form is the
most important and promising characteristic of these EOS.

19
Although all four families of EOS received considerable attention in the
past, recent literature mainly contains the developments in EOS of the vdW
family because of their simplicity and practicality. Tsonopoulos and Heidman^i
attribute the widespread use of cubic equations of state for describing PVT
relationships of oil/gas and phase equilibrium computations to the following
reasons:
Cubic EOS of van der Waals families yield relatively simple expression
for the thermodynamic properties and phase equilibrium relationships
of interest.
The EOS of the other three families, though more complicated in
nature, do not provide quantitatively better descriptions of phase
behavior than cubic EOS.
The intent of this section is to present a qualitative review of some widely
used cubic type equations of state, their applications, and tuning process of these
equations for different hydrocarbon systems.

2.3.1 Cubic Equations of State


Cubic equations of state are commonly used in the petroleum industry,
particularly in compositional reservoir simulators, to predict the phase behavior
of petroleum reservoir fluids. Since van der Waal proposed his EOS in 1873,
numerous modified versions with two or more parameters have been developed
to improve predictions of volumetric, thermal, physical and phase equilibrium
properties of the fluids. In this study, only the equations which are widely used
in the petroleum industry and a few promising EOS recently developed have
been reviewed.
The van der Waals or any other cubic equation of state can be expressed in
a following generalized form:

20
P "" p repulsive " p attractive (2.11))

van der Waals proposed the following equation.


RT a

where the repulsive pressure term (prepuisive) is represented by term


RT/(V-b) and the attractive pressure term (pattractive) is described by a/V^. The
parameters a and b are constants characterizing the molecular properties of the
individual components. The parameter 'b' represents the real or hard-sphere
volume of the molecules, and 'a' represents the intermolecular attraction. These
parameters are normally determined by imposing critical conditions. Other
parameters in the above equation are defined as follows:
p = system pressure, psia
T = system temperature, R
R = gas constant, 10.73 psia-ftV (lb-mole)(oR)
V = molar volume, ft^/mole.
Equation 2.11 is usually referred as the van der Waal two parameter cubic
equation of state. The term "two parameter" refers to parameters a and b. For
vdW equations, these parameters can be computed from the following
expressions.

a = Q, '- (2.12)
Pc

RZ
b=Q , ^ (2.13)
Pc

Where, pc = critical pressure, psia


Tc = critical temperature, R
Q = 0.421875
Q^ = 0.125.

21
The term cubic equation of state implies an equation which, if expanded,
would contain volume terms raised to first, second, and third power. Equation
2.11 can be expressed in the following cubic form in terms of volume.

RT_
V'- b+ F ' + -\V- =0 (2.14)
PJ \pj \pJ

Equation 2.11 can be expressed in a more practical form in terms of


compressibility factor Z by replacing the molar volume V in equation 2.14 with
ZRT/p.

Z' - (1 + B)Z^ +AZ-AB =0 (2.15)

where, A=- ^ (2.16)

hp
B=^ - (2.17)

Equation 2.15, when solved, yields one real root in the single phase region
and three real roots in the two phase region, where the largest root corresponds
to the compressibility factor of the vapor phase Z^, and the smallest positive root
corresponds to the compressibility factor of liquid phase Z^.
The van der Waals equation of state provides a good qualitative
description of the PVT behavior of substances in the liquid and vapor state. But
it is not considered accurate enough to be suitable for design purposes. Several
investigators developed new and more accurate equations of state that can be
used for facility design and reservoir simulation purposes. All equations of state
are generally developed for pure fluids, then extended to mixtures through the
use of mixing rules. These mixing rules are means of calculating mixture
parameters equivalent to those of pure components. Some of most popular EOS
are next discussed.

22
2.3.2 Redlich-Kwong EOS and its Modifications
Redlich and Kwong^^ demonstrated that by a simple adjustment, the
attractive pressure term (a/V^) in the van der Waal equation can considerably
improve the prediction of the volumetric and physical properties of the vapor
phase. These investigators replaced the attractive pressure term with a
generalized temperature dependent term. Their equation has the following
form:
RT a

where all the parameters are as defined for vdW EOS except parameters a
and b. These parameters can be calculated from the following expressions:
n2a-2.5
a = Q, ^ (2.19)
Pc

b=Q , ^ (2.20)
Pc

Where, Q^ =0.42747
Q^ =0.08664.
The Redlich-Kwong EOS can be expressed in the following cubic form:

Z'-Z'+{A-B-B')Z-AB =0 (2.21)

where, A= 2^.2.5 (^-^2)

5 =-^. (2-23)
RT ^ '
These authors extended the application of their equation to hydrocarbon
liquid mixtures by employing the following mixing rules:

, = Sz,/0.5 (2.24)
.'=1

23
K=T^,b, (2.25)
/=1

where,
n = number of components in the mixture
ai = parameter a for the i* component as defined by equation 2.19
bi = parameter b for the i* component as defined by equation 2.20
am = parameter 'a' for a mixture
bm = parameter 'b' for a mixture
Zi = mole fraction of component i in the liquid or vapor phase.
To calculate am and bm for a hydrocarbon gas mixture with a composition
yi, replace Xi with yi in equations 2.24 and 2.25.
Several modification of Redlick-Kwong EOS are reported in the literature.
Zudkevitch and Joffe^^ and Joffe et al.^s modified the RK EOS by assuming that
the Qa and Qb parameters were temperature dependent. These authors obtained
Qa and Qb for each pure substance at any temperature from saturated liquid
density data, and equalization of the saturated liquid and vapor phase fugacity.
The term fugacity can be defines as the vapor pressure modified to represent
correctly the escaping tendency of the molecules from one phase into the other.
The fugacity of a component can be expressed in the following mathematical
form:

f = p*EXP^ ^ '^ ^ (2.26)

where,
/ = fugacity, psia
p = pressure, psia
Z = compressibility factor.
The ratio of the fugacity to the pressure, i.e., / / p , is called the fugacity

coefficient, O.

24
Soave26 kept the RK EOS volume function and introduced a temperature
dependent function to modify the attraction parameter. This suggestion by
Soave has prompted an enormous increase in activity on the part of scientists
and engineers who are interested in the use of EOS to predict fluid properties.
The Soave-Redlich-Kwong (SRK) equation of state has the following form:
RT aa

Where the dimensionless factor a is a function of temperature and is


defined as:

c^ = [l + m(l-7;')]' (2.28)

Soave corrected the slope (m) against the acentric factor (co) by the
following generalized relationship:
m = 0.480 +1.5746; - 0.176^;' (2.29)
where, Tr = reduced temperature, T/Tc
CO = acentric factor of the substance.
For any pure component, the constants can be found from the following
expressions:

fl = Q, '- (2.30)
Pc
RT
b = n,^ (2.31)

where Qa and Qb are the SRK dimensionless pure component parameters


and are independent of temperature, pressure, composition, or particular
component. For the SRK equation, these parameters have the following values:
Qa = 0.42747
Qb = 0.08664.
SRK equation can be expressed in the following cubic form:

25
Z' -Z^ +{A-B-B')Z-AB =0 (2.32)

Where Z is the compressibility factor and A and B are defined by:

^ =( ^ (2.33)

bp
B = -:- (2.34)
RT ^ ^
Soave suggested the following mixing rules for the vapor-liquid
equilibrium calculations:
i^^X, = Z Z ^,^/(////) [KJ -1) (2.35)

K, = T{^A) (2-36)

where Zi is the mole fraction of the component 'i' in the liquid or vapor
phase, and parameter kij is an empirically determined correction factor called the
binary interaction coefficient, characterizing the binary formed by component i
and component j in the hydrocarbon mixture. These binary interaction
coefficients are used to model the intermolecular interaction between the binary
components and are dependent on the difference in the molecular size of the
components. Slot-Petersen^^ described the following properties of binary
interaction coefficients:
a. The interaction between hydrocarbon components increases as the
relative difference between their molecular weight increases.
b. Hydrocarbon components with the same molecular weight have a
binary interaction coefficient of zero.
c. The binary interaction coefficient matrix is symmetric, i.e., ki,j = kj,i
The fugacity coefficient, OiS of a component 'i' in the liquid phase in a
mixture is represented by the following expression proposed by Soave^^:

26
\
A(2V, B\
ln{<^^)='-K^Mz'-B)-iB V V K)
In 1 +
Z'J
(2.37)
m

Where, , = S ^\^,^.i(^.^.i) (l-^/y) (2.38)

^ = Z Z hh[^>^.i^i^) [^-^i] (2.39)


' ./

where. Of = fugacity coefficient of component 'i' in the liquid phase


Z^ = compressibility factor of the liquid phase.
Equation 2.37 can also be used to determine the fugacity coefficient of
component 'i' in the vapor phase. Of, by using the composition of the vapor
phase, yi, in calculating A, B, Z \ and other composition dependent terms.
Graboski and Daubert^^ suggested that no binary interaction coefficients
are needed for hydrocarbon systems. However, when non-hydrocarbons are
present in the mixture, binary interaction coefficients can substantially enhance
the volumetric and phase behavior predictions of the mixture. Vidal and
Daubert29 and Elliot and Daubert^^ described the theoretical background of the
interaction coefficients and the quadratic mixing rules of the Soave equation.
Elliot, et al., proposed the following correlations for calculating binary
interaction coefficients:
For N2 system
k^j = 0.107089+ 2.9776A:; (2.40)

For CO2 systems

A:,^ =0.08058-0.77215A:; -1.8407(/:,;)' (2.41)

For H2S systems


k,j =0.07654+ 0.01792 U ; (2.42)

For methane systems with compounds of ten hydrocarbon or more

27
k,. = 0.17985 + 2.6958^; +10.853(yt;)' (2.43)

For the above correlations.

(2.44)
" " 2e.e
' ./

where, f. = [a, ln(2)]7-^/ . (2.45)


The SRK equation of state is found to be capable of predicting vapor-
liquid equilibrium conditions (pressure, temperature, composition) for
hydrocarbon mixtures, but it does not provide very satisfactory results for liquid
compressibility predictions.
Peneloux, Rauzy, and Freze^^ introduced a volume correction parameter,
Ci, into the SRK EOS to minimize the errors in the saturated liquid volumes as
predicted by the SRK EOS. The third parameter does not change the vapor-
liquid equilibrium conditions determined by the SRK equation, but affects the
liquid and gas volumes by the following translation along the volume axis:
RT
c, = 0.40768(0.29441 - Z ^ ^ ) - (2.46)
r^ci

where ZRA is the Rackett compressibility factor and is a unique constant


for each compound. The values of ZRA are in general not much different from
those of the critical compressibility factors Zc. If their values are not available,
Peneloux et al., proposed the following correlation for calculating Ci:
T
c, = 4.43797878(0.00261 + 0.0928^;,)^^ (2.47)
r^ci

where coi is the acentric factor for the component i.

28
2.3.3 Peng-Robinson EOS and its Modifications
Peng and Robinson^ proposed a modified Redlich-Kwong equation of
state capable of predicting the liquid density as well as vapor pressure in order
to further improve vapor-liquid equilibrium predictions. Their equation has the
following form:
RT aa
p = yz^-v(v+b)+b{v-b) ^^ ^
where parameters a and b are defined as:

(RTY
a = 0.457235^^-^^ (2.49)
Pc
RT
b = 0.011796-. (2.50)
Pc

These investigators used the form proposed by Soave to calculate a


(equation 2.27) with the following expression for calculating m:
m = 0.3796 +1.54226^; - 0.26992a;'. (2.51)
This expression was later modified to improve predictions for heavier
components by the investigators to give the following relationship:
w = 0.3796+ 1.485(y-0.1644fi;' +0.0\661co\ (2.52)
The Peng-Robinson equation can be arranged in the following cubic
compressibility form:
Z' +{B + \)Z^ +{A-3B' -2B)Z-{AB-B'-B'). (2.53)

The fugacity coefficient of component i in a hydrocarbon liquid mixture is


calculated from the following equation:

*/(^'-0 ,.i.:. M ^ 2vi_A Z^+2.4145^


H<t>,^)=-^^T^Mz'-B)- 2.828435
In
Z'-0.4145>

(2.54)

29
Mixture parameters in equations 2.53 and 2.54 (A, B, bm, vj/i, \|/) are defined
previously for the SRK equation.
Stryjek and Vera^*- 35 observed that large deviations between experimental
and calculated vapor pressures as a function of reduced temperature given by
the PR equation. They further stated that the observed deviations are large at all
temperatures for compounds with large acentric factors and that the errors
increase rapidly at low reduced temperatures for all compounds. These
investigators proposed that a major improvement can be obtained with the
following expression for m:

m = w +m,(l + r ; ' ) ( 0 . 7 - T;) (2.55)

where Tr is the reduced temperature of the pure component, mi is an


adjustable parameter characteristic of each pure component, and mo is defined
by the following expression:

w =0.378893 + 1.48971536;-0.17131848a;' +0.0196554fi;\ (2.56)


For all components with a reduced temperature above 0.7, Stryjek and
Vera recommended setting mi = 0.
Jhaveri and Youngren^^ found in their application of the PR equation to
reservoir fluids that the error in the prediction of the gas phase compressibility
factors range from three to five percent and the error in the liquid density
predictions from six to twelve percent. Following the procedure proposed by
Peneloux et al.,^^ these authors introduced the volume correction parameter Ci to
the PR equation of state. This third parameter had the same units as the second
parameter bi of the unmodified PR equation and is defined by the following
relationship:
c,. = S,b, (2.57)

where Si is the dimensionless parameter and is known as shift parameter.


The authors then presented values of the shift parameter for the well defined

30
hydrocarbons. They also outlined the characterization procedure of the shift
parameter for the heptane plus fractions. These investigators concluded that a
significant improvement in the volumetric prediction can be obtained by the
modified PR equation.
The problem of characterizing the C7+ fractions in terms of their critical
properties and acentric factors has been long recognized in the petroleum
industry. Changing the characterization of C7+ fractions present in even small
amounts can have a profound effect on the PVT properties and the phase
equilibria of a hydrocarbon system as predicted by Peng and Robinson equation
of state. Recognizing this problem, Ahmed^^ devised an approach to determine
parameters a, b, and a from the readily measured physical properties of C7+
fractions (molecular weight and specific gravity). The author proposed the
following expression for calculating parameters a, b, and a of C7+ fractions:

'520' 0.5\
a = 1+ m 1- \ T ) (2.58)

where m is defined by the following expression:


f \
i^/rln. + a, MW^ + a, MW^.^^ + " '
m
a,+a,(MF/r)J MW^
(2.59)
2 ^^8

7ci+

f MW^ f MW^ ^ MW^' a^yC1+


a = a^ +^2 + a. + a. +
^ r ^ci. ^ r \:i. \ Y J n. ^n. (2.60)
2 ^8
+ 6rC7. +7rc:7. + -
/ C7+

For the above expressions:


MWc7+ = molecular weight of C7+
yc7+ = specific gravity of C7+

31
ai-as = coefficients as given by Ahmed.^^

2.3.4 Schmidt and Wenzel EOS


The PR and SRK equations of state predict a critical compressibility factor,
Zc, of 0.307 and 0.333, respectively, for all substances, whereas Zc is in the range
of 0.2 to 0.3 for most real substances. Schmidt and Wenzel (SW)^^ introduced a
component dependent compressibility factor by incorporating the acentric factor
in the van der Waal equation. The three parameter SW equation of state has the
following form:
RT aa
P ^ V^ ~ V'+(\ + 3co)bV-3a)b' ^^-^^^
with

i^'^cf , RT,.
a = Q^--^ and b = Q,-
re Pc

where, Q = ( l - 4 ( l - y 9 , ) ) ' (2.62)

a,=M.. (2.63)
The parameter Pc is given by the smallest positive root of the following
equation:
[6co + \)pl + 3/3^ + 3/?, - 1 = 0 (2.64)
and

4 = ,/, \ y (2.65)
3(1 + p^o))
Equation 2.64 can be solved for Pc by employing the Newton-Raphson
iterative method with a starting value of:
p^ = 0.25989 - 0.02176; + 0.00375^;'. (2.66)
The dimensionless factor a is a function of temperature and is defined by
following equations:

32
For Tr > 1, a = \- (0.4774 +1.328^;) ln(r,) (2.67)

ForTr<l, a = [\ + m{\-T;')]\ (2.68)


The parameter m in the above expression is correlated with the acentric
factor and the reduced temperature. The authors propose the following
expression for slope m:
w = w, For CO < 0.4
m = m2 For co > 0.5

m = [(0.55 - co)m^ +(o)- 0.40)w2 ] /0.15 For 0.4 < co < 0.5
with

m,=m,^+{yjQ){5T^-3m,-\f (2.69)

^2 = mo+0.71(7,-0.779)' (2.70)
where,
w = 0.465 +1.3476; - 0.528^' For co < 0.3671 (2.71)
w = 0.5361 + 0.9593a; For co > 0.3671. (2.72)

The Schmidt-Wenzel equation can be rearranged in cubic compressibility


factor terms to give the following form:
Z' +{UB-B-\)Z'' +(WB^ -UB^ UB + A)Z
~ (2 73)
-[WB' ^WB"- +AB) = 0
For hydrocarbon mixtures, the coefficients of equation 2.72 are defined by
the following equations:

A = ^->rr.j and B =
R^T^ RT
U = \ + 3cD, (2.74)
W = -3cu, (2.75)

33
Schmidt and Wenzel adopted the following mixing rules for calculating
the mixture parameters am, bm, and com:

. = Z Z ^/^/(,>) (l-^/y)
/=1 7=1 L

/=1

<",,,= (2.76)
lUft"'
The three parameter Schmidt-Wenzel equation has been reported to be as
capable as Peng-Robinson22 and Soave-Redlich-Kwong^^ equations in predicting
the vapor pressure while improving the predicted liquid density.

2.3.5 Patel and Tela EOS


Patel and Teja (PT)^^ developed a cubic equation of state by introducing
two substance dependent parameters which are obtained from the liquid density
and vapor pressure data. These authors corrected the two parameters with the
acentric factor. Patel-Teja proposed the following form of EOS:
RT _ aa
P-y-b v'+(b + c)V-bc ^ ^
Where parameters a, b, c, and a are defined by the following expressions:

fl = Q. (2.78)
Pc

RT^
b = Q, (2.79)
Pc

RT
c = Q, (2.80)
Pc

a= l + m(l-r)f (2.81)

34
where,

n = 34' + 3(1 - 2^ )a, + n^ + (i - 34) (2.82)


f^c = 1 - 34 (2.83)
and Qb is the smallest positive root of the following cubic equation:

Q^, + ( 2 - 3 i ) Q l + 34'Q, -i'=0. (2.84)

Equation 2.84 can be solved for Qb by employing the Newton-Raphson


iterative technique with an initial value of Qb is given by following equation:
Q^ = 0.32429Z,, - 0.0022005. (2.85)
The P-T equation requires the knowledge of four characterization constant
Tc, pc 5c, and m for any fluid of interest. The authors points out that ^c is not
equal to the experimental compressibility factor, but is obtained from one or
more liquid density data and m is obtained from the vapor pressure of the pure
fluids. They proposed the following generalized expressions for the parameters:
m = 0.452413 + 1.30982fi; - 0.295937^;' (2.86)
4 = 0.329032 - 0.07679926; + 0.0211947^;'. (2.87)
The P-T equation (equation 2.77) can be used for the calculations of the
mixture properties, if the constants a, b, and c are replaced by the mixture
constants am, bm, and Cm, respectively. The mixture constants are defined as
follows:

, = ZS[z-z,/(.,,.)'(l - *.y)l (2.88)

6,=Ez,6, (2.89)

e, Y.z.c,. (2.90)

Equation 2.77 can also be rearranged to produce the following cubic form
in terms of compressibility factor:

35
Z'+(C-I)Z'+{A-2BC-B-C-B')Z+{BC+B'C-AB) =O
(2.91)
where, for mixtures

^n,P
A= (2.92)
(RTf

5 = ^ (2.93)
RT

C = EEE (2.94)
RT
For the Patel-Teja equation of state, the fugacity coefflcient is deflned by
the following expression:

b, ( y^A, (Q + d\ A( b,+c, )
In
Z-B\RTdJ \Q-d) + 2KQ'-d'J +
ln(o,) = - l n ( Z - 5 ) + ^
RT Q + r^
d] 2Qd
-{cX3B + C) + bX3C + B)}
In
8 Q-dJ Q'-d'

(2.95)
where.

^ / = E ^./(./) (l-^v) (2.96)

e = Z +0.5(5 +C) (2.97)

d = [BC + 0.25(B + Cy]'\ (2.98)

Patel and Teja^^ concluded that the proposed EOS is capable of yielding
accurate vapor-liquid equilibrium predictions while reproducing liquid and
vapor phase densities with sufficient accuracy.
Willman and Teja^o proposed that for undefined components such as C7+
fractions, the parameters m and ^c in PT equation can be correlated in terms of
the boiling point Tb and specific gravity y. They proposed the following
relationships:

36
fl, 2 ^3

^ = K+bJ,+b,r + b,T,r' (2.100)


Valderrama and Cisternas^^ modified the PT equation by using the critical
compressibility factor instead of the parameter ^c to correlate parameters of PT
EOS. They proposed the following correlations:

Q = 0.69368018 - 1.0633424Z.. +0.6828995Z3 - 0.2104403Z,'


4 (2.101)
+ 0.003752658Z

Q, = 0.025987178 + 0.180754784Z,. + 0.06125 8949Z^^ (2.102)


Q, = 0.577500514-1.898414283Z^. (2.103)

m = -6.608 + 70.43Z,. -159.0Z,'. (2.104)


Valderrama^^ later introduced both the critical compressibility factor and
acentric factor as generalized parameters in the PT equation.

2.4 Applications of Equations of State


The equations of state are widely used in petroleum and chemical process
industries for process design and reservoir simulation purposes. Some of the
most common application of equations of state are:
Determination of equilibrium constants
Determination of dew point pressure
Determination of bubble point pressure
Determination of mixture critical properties
Two and three-phase equilibrium calculations.
Apart from above the applications, equations of state are also commonly
used in petroleum industry for simulating the PVT properties of the reservoir
fluids in place of costly experimental PVT tests. Some of experimental
procedures which can be adequately simulated using equations of state are:

37
Constant volume depletion test
Constant composition expansion test
Differential liberation test
Flash separation test
Composite liberation test
Swelling test.
Some of the applications of equations of state which are employed in this
study are next discussed in detail.

2.4.1 Determination of Equilibrium Constants


The system temperature (T), system pressure (p), and the overall
composition of the mixture must be known to determine the equilibrium
constant of a component in a mixture. The step-by-step procedure for
determining the equilibrium constants, as described by Ahmed,^^ is as follows:
Step 1: Assume an initial value of the equilibrium ratio for each
component in the mixture at the specified temperature and pressure. Wilson's.^^
equation can provide the starting values for equilibrium ratios.
5.37(1+^, )[l-^)
A Pa
K; = ^^ EXP (2.2)
p
where,
K,^ = assumed equilibrium ratio of component 'i'.
Step 2: Using the overall composition and assumed K-values, perform
flash calculations.
Step 3: Using the calculated composition of the liquid phase, Xi, determine
the fugacity coefficient, OiS for each component in the liquid phase from the
desired equation of state method.

38
Step 4: Repeat step three by using the calculated composition of the vapor
phase, yi, to determine the fugacity coefficient, Oi^, for each component in the
vapor phase.
Step 5: Calculate the new set of equilibrium ratios by employing the
following equation:

OL
K,=r- (2.105)
/

Step 6: Check for the solution by applying the following condition:

z
/=]
< G (2.106)

where,
e = preset error tolerance, e.g., 0.0001
n = number of components in the system.
If the above condition is satisfied, then the solution has been reached. If
not, steps one through six are repeated by using the calculated equilibrium ratios
as initial values.

2.4.2 Simulation of Constant Volume Depletion Test


The constant volume depletion (CVD) tests are performed on a gas -
condensate reservoir fluid sample in such a manner as to simulate pressure
depletion of an actual reservoir, assuming that retrograde liquid appearing
during production remains immobile in the reservoir. In the absence of the CVD
test data on a specific gas condensate system, the prediction of pressure
depletion behavior can be obtained by using any of the well established
equations of state to compute the phase behavior when the composition of the
total gas condensate system is known. The step by step procedure, as described
by Ahmed,^^ is summarized below:

39
Step 1: Assume that the original hydrocarbon system with a total
composition (zi) occupies an initial volume of one cubic feet at the dew point
pressure (pd) and system temperature (T):
Vi = l.
Step 2: Calculate the gas compressibility factor at dew point pressure (Zd)
from compressibility factor cubic equation for chosen equation of state, e.g.,
equation 2.53 for Peng-Robinson equation of state.
Step 3: Calculate the initial moles in place by applying the real gas law

" ' = ^ - P.107)

Step 4: Reduce the pressure to a predetermined value p. At this pressure


level, the equilibrium ratios are calculated and used in performing flash
calculations. The calculated numerical results include:
equilibrium ratios
composition of the liquid phase (retrograde liquid), Xi
moles of the liquid phase, UL
composition of the vapor phase, yi
moles of the gas phase, nv
compressibility factor of the liquid phase, Z^
compressibility factor of the vapor phase, Zy.
Step 5: Because the flash calculations are usually performed assuming the
total moles are equal to one, calculate the actual moles of the liquid and vapor
phases from:
(nL)actuai = ninL (2.108)
(nv)actuai = ninv. (2.109)
Step 6: Calculate the volume of each hydrocarbon phase by applying the
following expressions:

40
(nA Z'RT
V, = ^-^'^ (2.110)

(ny) Z'RT
V, = ^-^'^ (2.111)

where,
F^ = volume of the liquid phase
V^ = volume of the vapor phase.

Step 7: Calculate the total volume of the fluid in the cell


K = VL + V, (2.112)
where V, = total volume of the fluid, ft^.

Step 8: Since the volume of the cell is constant at one cubic feet, remove
the following excess gas volume from the cell:

Step 9: Calculate the number of moles of gas removed

^ = ^ F ^ . (2.114)

Step 10: Calculate cumulative gas produced as a percentage of gas

initially in place by dividing cumulative moles of gas removed ( X ^ p ) ^7 ^^e

original moles in place

%G, = '^- 100. (2.115)


( " - )',original

Step 11: Calculate the two-phase gas deviation factor from the following
relationship:

41
i^twophase / \ . (2.116)
[n^-n^)RT
Step 12: Calculate the remaining moles of gas by subtracting the moles
produced from the actual number of moles of the vapor phase

M.-M.,^-".- (2.117)
Step 13: Calculate the new total moles and new composition remaining in
the cell by applying the molal and component balances

"={"X..A"A (2.118)

z, = '^ "'-'-' '^''' . (2.119)


;?.

Step 14: Consider a new lower pressure and repeat steps four through
thirteen.

2.4.3 Simulation of Constant Composition Expansion Test


Constant composition expansion experiments, commonly called
pressure-volume tests, are performed on gas condensate or crude oil to simulate
the pressure-volume relations of these hydrocarbon systems. These tests are
conducted to determine:
saturation pressure (bubble point or dew point pressure)
compressibility factor of the gas phase
total hydrocarbon volume as a function of pressure.
The simulation procedure of constant composition expansion test, as
described by Ahmed,!^ is illustrated in the following steps:
Step 1: Given the total composition of the hydrocarbon system, Zi, and
saturation pressure (bubble point pressure, pb, for oil systems and dew point

42
pressure, pd, for gas systems), calculate the total volume occupied by one mole of
the system at saturation pressure by applying the following expression:
{\)ZRT
K. = - ^ (2.120)

where,
Kai ~ volume at saturation pressure, ft^/mole
p^.^, = saturation pressure, psia
Z = compressibility factor Z^ or Z^ depending on the type of system.
Step 2: The pressure is increased in steps above the saturation pressure.
At each pressure, the compressibility factor Z^ or Z^ is calculated from
compressibility factor cubic equation for the chosen equation of state, e.g.,
equation 2.53 for PR EOS, and used to determine the fluid volume
^. _ JDZRT
P
where,
V = compressed liquid or gas volume at the pressure p, ft^/mole
Z = compressibility factor of the compressed liquid or gas.
The corresponding relative phase volume (Vrei) is calculated from the
expression:
V
V =
' sal

Step 3: The pressure is then reduced in steps below the saturation


pressure. The equilibrium ratios are calculated and flash calculations are
performed at each pressure level. The resulting data include Ki, Xi, yi, nL, nv, ZK
and Z^. Since no hydrocarbon material is removed during pressure depletion,
the original moles (ni=l) and composition remain constant. The volumes of
liquid and gas phases can then be calculated from the following expressions:

43
V, = - ^ (2.121)

V, = - ^ ^ (2.122)

and

where, Vt = total volume of the hydrocarbon system.


Step 4: Calculate the relative total volume from the expression:
V
Relative total volume = ^.
'^ sal

2.4.4 Simulation of Differential Liberation Test


The simulation procedure, as described by Ahmed,^^ jg summarized in the
following steps:
Step 1: Starting with saturation pressure (psat) and reservoir temperature
(T), calculate the volume occupied by one mole of the hydrocarbon system with a
overall composition of Zi by applying equation 2.120.
Step 2: Reduce the pressure to a predetermined value at which the
equilibrium ratios are calculated and used in performing flash calculations. The
actual number of moles of liquid phase with a composition of Xi, actual number
of moles of the vapor phase with a composition of yi, are then calculated from the
expressions:
(nL)actuai = ninL (2.108)
(nv)actuai = ninv. (2.109)
Step 3: Determine the volume of liquid and gas from following
expressions:

44
(nA Z'RT
y ^ ^ \ ^)aclual P^^Qj

(ny) Z'RT
y \Jacluai P^^^.
P
The volume of the produced gas is determined from the following
relationship:

G,=379.4(,)_, (2.123)

where, G^, = gas produced, scf.

The total gas produced can also be calculated from the expression:

Step 4: Assume that all the equilibrium gas is removed from contact with
the oil. This can be mathematically achieved by setting the overall composition,
Zi, equal to the composition of the liquid phase, Xi, and setting the total moles
equal to the moles of the liquid phase.
Zi = X i

n i = (nL)actual

Step 5: Using the new overall composition and total moles, step two
through step four are repeated. When the depletion reaches the atmospheric
pressure, the temperature is changed to 60 ^F and the residual oil volume is
calculated.
Step 6: The calculated volumes of the oil and removed gas are then
divided by the residual oil volume to calculate the relative oil volume (Bod) and
the solution gas-oil ratio (Rsd) from the following expressions:
Bod = VL/Vsat (2.124)
Rsd = (5.615)(volume of gas in solution)/(Vsat). (2.125)

45
2.5 Comparison of Equations of State
Numerous equations of state have been proposed and applied for phase
behavior modelling and compositional reservoir simulations. The more
commonly used equations have been compared by a number of
investigators,!-^' ^^ resulting mainly in a general conclusion that none of the
available equations can be singled out as the most superior one to predict all the
properties at all conditions. A number of comparative studies^- ^ limited to
predictions of volumetric and vapor-liquid equilibrium properties of petroleum
reservoir fluids have shown that certain equations exhibit a higher overall
accuracy. But the reliability of these equations varies for different fluid systems
and conditions.
Danesh et al.^'^^ studied extensively the reliability of ten leading equations
of state for predicting the phase behavior and volumetric properties of
hydrocarbon fluids, with particular relevance to North Sea gas injection systems.
Based on the experimental and simulation results, the authors found the
following conclusions:
1. The Patel-Teja equation with modification by Valderrama and Redlich-
Kwong equation with modification by Zudkevitch and Joffe are
generally superior than other equations.
2. All the equations of state exhibit a similar convergence behavior. They
either could not easily converge or converged to the trivial solution of
equal partitioning of the components between the phases at conditions
close to the predicted critical point.
Ahmed^ conducted a study to evaluate eight commonly used equations of
state to compare their ability to predict the volumetric and phase equilibria of
gas condensate systems and offered the following conclusions:

46
1. The Schmidt-Wenzel equation of state exhibits a superior predictive
capability for volumetric properties of gas condensate systems.
2. The Patel-Teja and Schmidt-Wenzel equations are found to give
reliable gas compressibility predictions.
3. In terms of vapor-liquid equilibrium predictions, the Peng-Robinson,
the Patel-Teja, and Schmidt-Wenzel equations all perform equally well.

2.6 Tuning of Equation of State


Cubic equations of state have found widespread acceptance as tools which
permit the convenient and flexible calculations of the phase behavior of the
reservoir fluids. They facilitate calculations of complex behavior associated with
rich condensates, volatile oils, and gas injection processes. Despite their
flexibility, the parameters of a cubic equation of state often need adjusting prior
to application to a particular reservoir fluid. For an EOS to reproduce
experimental data, it is necessary to adjust some of its parameters.^^' ^^ This
adjustment is necessary since the existing equations of state are only
approximations when applied to multi-component systems, besides the presence
of ill-defined mixtures of hydrocarbons, known as plus fractions, for which
information relevant to equations of state is usually not available. Adjusting the
parameters to overcome these limitations is called tuning or characterizing an
equation of state.
The parameters required for employing an equation of state are:
1. The critical temperatures, Tc, the critical pressures, pc, and the acentric
factor, CO, of each component of the mixtures.
2. The binary interaction coefficients, kij, between each pair of
components.
3. The molar composition.

47
For the well defined components of the mixture (N2, CO2, Ci-Ce), accurate
measured data are normally available for Tc, pc and acentric factor, and it is
usually not considered justified to use these as tuning parameters. However,
critical pressures and temperatures, and acentric factors for plus fractions are
usually determined from empirical correlation's which are prone to errors.
Therefore, it is considered justified to use Tc, pc, and co of the plus fractions as
tuning parameters.
There are two common approaches by which the equations of state are
tuned. One approach, proposed by Whitson,^^ consists of keeping the theoretical
parameters of the equation and performing the match on the critical properties
(Tc and pc) and the acentric factor of the plus fractions. The second approach,
proposed by Coats and Smart,^^ consists of calculating the critical properties and
acentric factor of plus fractions from available correlations and adjusting
theoretical equation of state parameters Qa and Qb for methane and plus fraction
and the binary interaction fractions between methane and plus fractions by using
non-linear regression to laboratory data. The laboratory data might include
saturation pressures, densities, equilibrium constants, constant volume
expansions, differential liberation, etc. Merril et al.^^ discussed in detail about
the sort of data that should be used in the tuning process. Pedersen et al.^i
recommended the use of molecular weight of plus fraction as tuning parameter.
These authors argued that the experimental uncertainty on the molecular weight
of the plus fraction is normally of the order of five to ten percent and, therefore,
justified to adjust the molecular weight of the plus fraction.
It is generally accepted that proper tuning of an equation of state can
substantially enhance the accuracy of volumetric and phase behavior predictions.
Pedersen et al.,^^ however, warned that the indiscriminate tuning can lead to
inconsistent predictions. The authors further stressed that the tuning procedure

48
will not correct the deficiencies of the equation of state used and that the EOS
predictive capabilities depend entirely on the type and the accuracy of the data
used for the tuning. The investigators emphasized that attempts should be made
to ensure that the tuned parameters are within reasonable physical limits.

2.7 Characterization of Plus Fractions


The hydrocarbon plus fractions that comprise a significant portion of
naturally occurring hydrocarbon fluids create major problems when predicting
the thermodynamic properties and the volumetric behavior of these fluids by
equations of state. These problems arise due to the difficulty of properly
characterizing the plus fractions in terms of their critical properties and acentric
factors.
Whitson^^ have shown the effect of the heavy fractions characterization
procedure on PVT relationships prediction by equation of state. Usually, these
undefined plus fractions, commonly known as C7+ fractions, contain large
number of components with a carbon number greater than six. Molecular
weight and specific gravity of the C7+ fractions may be the only measured data
available. In the absence of detailed analytical data for the plus fraction in a
hydrocarbon mixture, erroneous predictions and conclusions can result if the
plus fraction is used directly as a single component in the mixture phase
behavior calculations. Hariu et al.^^ suggested that these errors can be
substantially reduced by splitting or breaking down the plus fraction into a
manageable number of fractions (pseudo-components) for equation of state
calculations.
Splitting or breaking down schemes refer to the procedures of dividing
the heptanes-plus fraction into hydrocarbon groups with a single carbon number
(C7, Cs, C9, etc.) and are described by the same physical properties used for pure

49
components. Several investigators^^' ^^-^^ proposed different schemes for
extending the molar distribution behavior of C7+, i.e., the molecular weight and
specific gravity. In general, the proposed schemes are based on the observation
that lighter systems such as condensates usually exhibit exponential molar
distribution, while heavier systems often show left-skewed behavior. This
behavior is shown schematically in Figure 2.8.
Three important requirement should be satisfied when applying any of
the proposed splitting models. These requirement are:
the sum of the mole fractions of the individual pseudo-components is
equal to the mole fraction of C7+.
the sum of the products of the mole fraction and the molecular weight
of the individual pseudo-components is equal to the product of mole
fraction and molecular weight of C7+.
the sum of the product of the mole fraction and molecular weight
divided by the specific gravity of each individual component is equal
to that of C7+.

50
LU
CC

(f)
LU
DC
QL

SPECIFIC VOLUME

Figure 2.1
P-V-T Diagram for a Single-Component System
(After Reference 12, Page 80)

51
P
r
e
s
s
u
r
e

Temperature

Figure 2.2
Pressure-Temperature Diagram for a Single-Component System

52
o
CO
</)
0)

Pd = Pt <

I
JL.
Vc

Volume

Figure 2.3
Pressure-Volume Diagram for a Single-Component System
(After Reference 14, Page 3)

53
LI qu \ii

a.

Gas

luKipcr <i l u r e

Figure 2.4
Typical Pressure-Temperature Diagram for a Binary System
(After Reference 14, Page 17)

54
Temperature

Figure 2.5
Effect of Composition on Phase Envelope for a Binary System
(After Reference 12, Page 85)

55
Temperature

Figure 2.6
Location of Critical Loci for Several Binary Systems
(After Reference 12, Page 85)

56
l i i K l e r s a t u r a t e d Oil Reservoir

H r 1
Ci n u ij 1 I'oiiiL

b Cricondenbar

<-^x _ _CLiLical_ Prgs^ur^e _ __ j^j

Saturated
Reservoir

E Cricondentherm

Two-Phase /
Region

ijc ^ ct
Temperature

Figure 2.7
Typical Pressure-Temperature Diagram for a Multi-Component System
(After Reference 14, Page 22)

57
M
o
1 Left-Skewed Distribution
e (Heavy Hydrocarbon System)

F
r
a
c
t
I
o
n

Exponential Distribution
(Condensate System)

Molecular Weight

Figure 2.8
The Exponential and Left-Skewed Distribution Functions

58
CHAPTER 3
METHODOLOGY

The experimental data available from the PVT studies of two hydrocarbon
fluids were used to characterize the vaporization/condensation process of a low
density crude oil by elevating the temperature at constant pressure. The
experimental data included the composition of the hydrocarbon fluid at the
initial reservoir conditions and data from the constant composition expansion
test and constant volume depletion test.
Initially, an attempt was made to characterize the entire
vaporization/condensation process on HYSIM,^^ a process simulator provided
by Hypertech Limited. Although, this program provided good results for K-
value predictions and flash calculations, it was not capable of simulating the PVT
experimental procedures such as constant volume depletion test and constant
composition test which were required in this study.
Therefore, the entire characterization process was performed on WinProp,
a phase property program provided by Computer Modelling Group.^'^ This
program was capable of predicting K-values and performing flash calculations as
well as simulating required laboratory procedures.
This study was conducted in the following stages to achieve the study
objectives:
1. Compositional data was set for both fluids at initial pressure and
temperature conditions.
2. The most suitable equation of state model was identified by simulating
the constant composition expansion and constant volume depletion
tests using three equation of state models and comparing the
simulation results with the available experimental data.

59
3. The selected equation of state model was employed to describe
vaporization/condensation process for a low pressure, hydrocarbon
liquid-vapor system by elevating the temperature.
The following sections provide the detailed description of the procedure
during each of the above mentioned stages of the study. The adopted procedure
is also shown schematically in figure 3.1.

3.1 Setting-Up the Initial Data and Characterization of


Plus Fractions
The composition of the fluids, referred as sample one and sample two,
were specified along with the initial pressure and temperature conditions. The
original composition of both the samples is listed in table 3.1. The initial
conditions for sample one and two were 5707 psig, 264 F and 5713 psig, 186 F,
respectively. In this stage, the following properties for each component were
specified for later use in the equation of state:
Molecular Weight (MW)
Critical Pressure (pc)
Critical Temperature (Tc)
Acentric Factor (co)
Specific Gravity (SG)
Normal Boiling Point Temperature (Tb).
For the well-defined pure components such as CO2, N2, Ci-Ce, all of these
properties are very well researched and widely reported in the literature.
However, this is not the case with an undefined mixture of components such as
hydrocarbon plus fractions or pseudo-components. Therefore, to characterize
the plus fractions, these fractions were divided into single carbon number (SCN)
fractions based on the molecular weight, specific gravity of the plus fractions
(table 3.2) and the fluid type. Since, it was known from the PVT experimental

60
results that both the fluids were retrograde condensate type, an exponential
molar distribution profile was specified for splitting hydrocarbon plus fractions
into single carbon number fractions. This was achieved by employing the built-
in splitting scheme of the software WinProp. This scheme estimates internally
the number of pseudo-components (SCN fractions) for the plus fractions. The
above mentioned properties of these pseudo-components were calculated from
the empirical correlations suggested by Lee and Kesler.20

3.2 Selection of Equation of State Model


The objective of this stage was to tune various equations of state to match
available experimental data and select the equation of state which produces the
closest match. Once the composition and properties of each pure and pseudo-
component were set, the following three equation of state models were used
separately to simulate the constant composition expansion test and constant
volume depletion test:
1. Peng-Robinson equation of state (PREOS).
2. Original Soave-Redlich-Kwong equation of state (SRKEOS).
3. Soave-Redlich-Kwong equation of state with the constant 'a' proposed
by Graboski and Daubert (SRKGDEOS).
For an equation of state to accurately reproduce the experimental data, it
was necessary to tune some of the equation of state parameters. This was
achieved by employing the regression procedure suggested by Agarwal et al.^^ to
tune the above mentioned equations of state. This procedure orders the
parameters, from the specified list of regression parameters, in such a way that
the most sensitive parameters are used first. The program "WinProp" has the
built-in feature that determines internally the most sensitive parameters. The
regression is performed by regressing on a small number of parameters at a time

61
(five in this study). Once a parameter reaches the maximum or minimum
allowed or does not contribute any longer to improve the match, it is replaced by
the next parameter that has not been used from the ordered list. Thus, it is
possible to specify a large set of regression parameters, and this procedure will
regress on a small number of parameters at a time, working from the most
sensitive parameters to least sensitive parameters.
The regression parameters used in this study were:
Critical pressures of pseudo-components (SCN fractions).
Critical temperatures of pseudo-components.
Acentric factors of pseudo-components.
Molecular weight of pseudo-components.
Volume shift parameter of methane and pseudo-components.
Binary interaction coefficient between CO2 and pseudo-components.
Binary interaction coefficient between methane and pseudo-
components.
Since the component properties, such as critical pressure, critical
temperature, acentric factor, and molecular weight of pseudo-components were
calculated from empirical correlations proposed by Lee and Kesler^o w^hich has a
error range o f + / - 1 0 % , a regression range of +/-10% was specified to tune these
parameters. For binary interaction coefficients, a trial and error method was
adopted to adjust these coefficients.
The experimental data used to tune the equations of state were relative
volume and equilibrium gas deviation factor data from constant composition
expansion test and retrograde liquid volume (% of hydrocarbon pore volume)
and produced gas (% of original moles) data from the constant volume depletion
test. The experimental data for samples one and two are presented in Tables 3.3
and 3.4, respectively.

62
Although, the prediction accuracy of the three equations of state models
were compared for all the above mentioned experimental data, the selection of
the equation of state model was based entirely on the accuracy of a particular
equation of state model in predicting the constant volume depletion test data
(retrograde liquid volume, produced gas). This was done because the simulation
of constant volume depletion test was used extensively in characterizing the
vaporization/condensation process in this study.

3.3 Characterization of Vaporization/Condensation


Process
The characterization of the vaporization/condensation process was
conducted at a low depleted pressure using the selected equation of state model.
Since, it was necessary to know the reservoir fluid composition at the reduced
pressure for the characterization process, a simulation of a constant volume
depletion test was conducted on the original hydrocarbon fluid to calculate fluid
compositions at specified depletion pressures (750 psia and 250 psia). These
pressures were constant throughout the vaporization/condensation process.
To check the accuracy of the calculated compositions, the CVD test was
also simulated at depletion pressures at which experimentally obtained
compositions were available. The calculated compositions were then compared
with experimental compositions, and it was assumed that the same amount of
error would be present in the composition calculated at a specified depletion
pressure. The experimental compositions were available at 715 psia for sample 1
and 619.7 psia for sample 2. It was assumed that reservoir temperature would
remain constant throughout the depletion process.
The vaporization/condensation process is shown schematically in Figure
3.2.

63
Once, the composition at the specified depletion pressure was available,
the following step-by-step procedure was adopted to characterize the
vaporization/ condensation process:
1. Flash calculations were performed on the calculated composition at the
specified depletion pressure and reservoir temperature, assuming 100
moles of fluid in place, to obtain the mole percent of liquid and vapor,
volume percent of liquid and vapor, liquid composition and vapor
composition.
2. The vaporization process was characterized by simulating the CVD
test at the depletion pressure and specified elevated temperature. The
elevated temperatures used were 700 F and 500 F for sample one and
500 F and 350 F for sample two. Different temperatures were used
for samples one and two because it was found that the fluid was
vaporizing completely at temperatures above 550 F in case of sample
two. In the laboratory CVD test, the sample of reservoir fluid is placed
in the test cell at the original pressure and temperature, and the
pressure in the cell is brought down to a specified pressure. The
reduction of pressure results in the generation of additional gas
(produced gas) which is expelled from the test cell to a separator. The
simulation of a CVD test was considered ideal for characterizing the
vaporization process in the reservoir because both generate excess gas
which is removed from the place where this gas was generated. The
difference between the CVD test and vaporization process is that in a
CVD test, the excess gases are generated due to reduction in pressure,
while, in the vaporization process, the excess gases are generated due
to an increase in temperature. The other difference is that in the CVD
test, the produced gas is sent to a separator, while, in the case of

64
vaporization, this produced gas is recombined with the depleted liquid
volume and composition at the original temperature. The output
obtained from the simulation of the CVD test were the mole percent,
and the compositions of the produced gas and liquid left in the test
cell.
3. The condensation process was characterized by moving the produced
gas from the vaporization stage to the PVT cell which is at the original
reservoir temperature and equilibrating the produced gas with the
liquid in place at the depleted pressure. It was assumed that the gas
moved from the vaporization stage displaces completely the gas in
place at the initial pressure and temperature. The overall composition
of the equilibrium liquid and vapor (produced gas and liquid in place)
was calculated by employing the following expression:

s '/produced \ '^ in-place ,_ ^v

^'" U]^x~)
V ^ ' nroduced s i'/in-place

where.
z, = mole fraction of component "i" in the mixture

(^y.) = mole fraction of component "i" in the produced gas

(jc ) = mole fraction of component "i" in the liquid in place


V '/in-place *

(^ny) = moles of produced gas

(n,) = moles of liquid in place.


V '^/in-place *

4. Flash calculation were performed on the new composition at the


depleted pressure to obtain mole percent and volume percent of the
liquid and vapor, respectively, and liquid and vapor compositions. At
this stage, the first vaporization/condensation cycle was completed.

65
5. After every vaporization/condensation cycle, the following condition
was tested:
[(Mole % of Liquid)n - (Mole % of Liquid)r,-i] < 1.0%
where,
(Mole % of Liquid)n = mole percent of liquid after "n" cycles
(Mole % of Liquid)n.i = mole percent of liquid after "n-1" cycles
This condition was applied because it was decided that the
vaporization/condensation process would be ineffective further if the
difference between the mole percent of liquid calculated after
consecutive cycles is less than one.
6. If the above condition was not satisfied, the whole process was
repeated beginning with the characterization of the vaporization
process using the new composition.
7. If the condition mentioned in step five was satisfied, then the possible
trends were observed by plotting separately mole percent of liquid and
vapor, volume percent (saturation) of liquid and vapor, overall
composition, liquid composition and vapor composition versus cycles.
Here one cycle is referred as one vaporization and condensation
process.
Once the characterization of the vaporization/condensation process for
one combination of reservoir pressure and elevated temperature was completed,
the whole process was repeated for other combinations of reservoir pressures
and elevated temperatures for both the samples and possible trends were noted.

3.4 Assumptions
The following summarizes the assumptions that were employed to
simplify the study and achieve the study objective:

66
1. To establish the fluid composition at depleted pressures, the reservoir
temperature was assumed constant during the depletion process.
2. The pressure would remain constant during the whole
vaporization/condensation process.
3. Composition and saturation of liquid and vapor phases were the same
at the initial and depleted reservoir conditions.
4. Residual liquid that appears during production would remain
immobile and could only be "produced" via vaporization.
5. A step temperature profile was assumed, i.e., the temperature
increases sharply from the initial temperature to the vaporization
temperature at the vaporization stage and decreases sharply from
vaporization temperature to the initial temperature at the
condensation stage.
6. Excess gas generated during the vaporization stage (produced gases in
the simulation of constant volume depletion tests), when placed with
the original liquid still at the original temperature, displaces the
original gas-in-place completely, i.e., total volume of produced gas
displaces completely the total volume of gas originally in place (at
initial temperature). This was accomplished by displacing moles of
gas in place with moles of produced gas.
7. The temperature increase at the vaporization stage does not result in
any change in overall composition, i.e., thermal cracking or burning
does not occur.

67
Table 3.1
Original Composition of Samples One and Two

Component Composition in Mole %


Sample 1 Sample 2
H2S Nil Nil
CO2 2.27 0.18
N2 0.72 0.13
Ci 69.01 61.92
C2 7.69 14.08
C3 3.71 8.35
i-a 1.59 0.97
n-C4 1.78 3.41
i-Cs 1.26 0.84
n-Cs 0.75 1.48
C6 1.72 1.79
C7+ 9.50 6.85

Table 3.2
Properties of Plus Fractions

C7+ Property Sample 1 Sample 2


Molecular Weight 174 143
Specific Gravity 0.786 0.795

68
Table 3.3
Experimental Data for Sample 1

Pressure Relative Equilibrium Retrograde Liquid Produced Gas


(psia) Volume Z-Factor Volume (% of Original
(% of HC Pore Volume) Moles)
7015 0.8955 - - -

6015 0.9396 - - -

5515 0.9696 - - -

5115 0.9989 - - -

5102 1.000 1.061 0.0 0.0


5098 1.0004 - 0.6 -

5075 1.0025 - 3.5 -

5015 1.0086 - 9.2 -

4918 1.0184 - 14.7 -

4415 1.0800 0.948 26.0 6.955


3615 1.2304 0.904 30.8 18.709
2815 1.5016 - - -

2715 - 0.894 30.6 34.922

2015 2.0501 - - -

1815 - 0.918 27.7 53.555

1637 2.5256 - - -

1402 2.9617 - - -

1215 3.4668 0.941 24.9 66.298

715 - 0.964 22.6 77.132

15 - - 18.5 -

69
Table 3.4
Experimental Data for Sample 2

Pressure Relative Equilibrium Retrograde Liquid Produced Gas


(psia) Volume Z-Factor Volume (% of Original
(% of HC Pore Volume) Mole)
6014.7 0.8808 - - -

5727.7 0.8948 1.107 - -

5314.7 0.9158 - - -

5014.7 0.9317 - - -

4414.7 0.9690 - - -

4014.7 1.000 0.867 0.0 0.0


3814.7 1.0194 - - -

3724.7 1.0299 - - -

3514.7 1.0559 0.799 3.3 5.374


3314.7 1.0878 - - -

3014.7 1.1496 - - -

2914.7 - 0.748 19.4 15.438


2719.7 1.2430 - - -

2219.7 1.5246 - - -

2114.7 - 0.762 23.9 35.096


1619.7 2.1035 - - -

1314.7 - 0.819 22.5 57.695


1024.7 3.5665 - - -

619.7 - 0.902 18.1 76.787


14.7 - - 12.6 93.515

70
Set-Up Initial Data
(Composition, Component Properties)

Split Plus Fractions Into SCN Fractions

Select Equation of State Model

Calculate Fluid Composition at Depleted Pressures

Perform Flash Calculations at Depleted Pressure and Reservoir


Temperature to Calculate Xi, yi. Mole % of liquid and Vapor,
Volume % of Liquid and Vapor

Characterize Vaporization Process By Simulating CVD Test


at Depleted Pressure and Vaporization Temperature

Recombine Produced Gas with Original Liquid at


Initial Conditions

Calculate New Composition by Adding Moles of


Produced Gas to the Moles of Liquid Initially in Place

Perform Flash Calculation on New Composition at Depleted Pressure and


Reservoir Temperature to Calculate New Xi, yi. Mole % of liquid and
Vapor, Volume % of Liquid and Vapor

[(Mole % of Liquid)new - (Mole % of Liquid)oid] < 1.0


If, No
If, Yes
Plot Cycles Versus Mole % and Volume % of Liquid
and Vapor, Overall Composition, Liquid
Composition, Vapor Composition

Figure 3.1
Flow Diagram of Procedure Used in the Study

71
o JJ
o
p.
en
O
5 "^ (N U
O ^
u C O

X c ... CD

u CD
/ :3 ^
u CD
O CD
u a O)
u >
o
;-!
O o CD rr, PH

CI. 03 CI. o
R> cs N
O o
.4-1

> ^ 2 ^ " CD

>
c;
-d
o
.11 o
C
.2 - I( >^
U
-d U
o -d CO o
CC
U o O) N
u
411 u O
O ^ I
U (-0
.O >

JJ

U D
CD o
.1t
.4-1

(-0

PH
a .- :i3 o
^ 1^ ^
vo

O II |_H

72
CHAPTER 4
DISCUSSION OF RESULTS

The identification of the most appropriate equation of state model and


characterization of vaporization/condensation process in the reservoir were
accomplished.
The equation of state model was identified by simulating a constant
volume depletion test and constant composition expansion test using three
equation of state models. The simulation results from these equations of state
were compared with experimental data, and the equation of state model that
predicted constant volume depletion test data most accurately was selected. The
characterization of the vaporization/condensation process in the reservoir was
accomplished through a combination of constant volume depletion test
simulation, and flash calculations using the selected equation of state model.
The following sections provide detailed description of results from
various stages of study.

4.1 Plus Fraction Characterization Results


The plus fraction of the reservoir fluid from both samples were split into
single carbon number (SCN) fractions and the SCN fraction properties were
calculated from empirical correlations proposed by Lee and Kesler.^i Plus
fractions from Sample one were split into five SCN fraction while plus fractions
from sample two were divided into four SCN fractions. The composition and
properties of SCN fractions from samples one and two are listed in Tables 4.1
and 4.2, respectively.

73
4.2 Selection of Equation of State Model
Three equation of state models were tested for accuracy in predicting
relative volume and equilibrium gas deviation factor from constant composition
expansion test simulation, and retrograde liquid volume and produced gas from
constant volume depletion test simulation.
It was found that all the equation of state models predicted relative
volume quite accurately (Figures 4.1 to 4.6) with the exception of SRKEOS and
SRKGDEOS in case of sample one (Figures 4.2 and 4.3). Among the three
models, the Peng-Robinson equation of state model was found to be most
accurate in predicting relative volume data (Figures 4.1 and 4.4). It was also
observed that none of the equation of state models were able to predict
equilibrium gas deviation factor (Figures 4.7 to 4.12) very well.
Although the accurate prediction of relative volume and gas deviation
factor were not the criterion for selection of the equation of state model, these
factors were calculated to verify the accuracy of the equation of state. The
selection of the equation of state was made on the basis of their accuracy in
predicting retrograde liquid volume and produced gas data. The results from
tested equation of state models in predicting the retrograde Hquid volume for
both sample are presented in Tables 4.3 and 4.4 and Figures 4.13 to 4.18 along
with the experimentally obtained data. It was observed from these tables and
figures that the values obtained from Peng-Robinson equation of state were the
closest to the experimental values. The results from tested the equation of state
models in predicting the produced gas for both sample are presented in Tables
4.5 and 4.6 and Figures 4.13 to 4.18 and again the Peng-Robinson equation of
state was found to be most accurate. Large errors were observed in predicting
retrograde liquid volume and produced gas from SRKEOS and SRKGDEOS in
case of sample one. These large errors were occurring probably due to inability

74
of these equation of state models in predicting accurate saturation pressures.
The error in calculating saturation pressure by these equations of state were
found to be large (Table 4.7) in case of sample one.
Based on the above mentioned results, the Peng-Robinson equation of
state model was selected to be used further in the study.

4.3 Results from Characterization of Vaporization/


Condensation Process
The first step in characterizing the vaporization/condensation process
was to calculate fluid composition at specified depleted conditions. This was
accomplished by simulating constant volume depletion tests on the original
hydrocarbon fluid using the Peng-Robinson equation of state. Fluid
compositions were also calculated at depletion pressures at which
experimentally obtained compositions were available. The comparison between
calculated composition and experimental composition was presented in Table 4.8
for sample one and Table 4.9 for sample two. The calculated composition for
both samples was found to be very accurate, and it was assumed that
composition calculated at specified depletion pressures had the same degree of
accuracy. The calculated composition at 750 psia and 250 psia are listed in Tables
4.10 and 4.11 for samples one and two, respectively.
The vaporization/condensation process was characterized at each selected
reservoir pressure (750 psia and 250 psia) and vaporization temperature (700 F
and 500 F for sample one, 500 F and 350 F for sample two) combination. The
following were the results from the characterization process:
Variation in mole percent of liquid and vapor with every
vaporization/condensation cycle was plotted for every pressure-
vaporization temperature combination for both samples (Figures 4.19
to 4.26). The mole percent of the liquid was found to be increasing

75
with every cycle, while the mole percent of the vapor was decreasing
with cycles in all the cases. However, the extent of this trend was
different for different combination of reservoir pressure and
vaporization temperature for both samples. This trend was significant
in the cases of a relatively high depleted pressure and high
vaporization temperature (Figures 4.19 and 4.23). This trend was less
significant in the cases of either low vaporization temperature (Figures
4.20 and 4.24) or lower depleted pressure (Figures 4.21 and 4.25) and
almost non-existent in the cases of low pressure and low temperature
(Figures 4.22 and 4.26). In the case of sample one, the mole percent of
liquid was increased from 44% to 96% at high pressure and high
temperature (Figure 4.19), while at the same pressure and low
temperature the mole percent of the liquid increased from 44% to 81%
(Figure 4.20). In the case of sample two, the mole percent of liquid was
increased from 44% to 93% at high pressure and high temperature
(Figure 4.23), while at the same pressure and low temperature the mole
percent of liquid changed from 44% to 11% (Figure 4.24). This
indicates that the vaporization/condensation process was more
effective at high vaporization temperature. The mole percent of liquid
was increased from 64% to 95% at low pressure and high temperature
in case of sample one (Figure 4.21), and from 64% to 93% in case of
sample two (Figure 4.25). This increase was significant but still less
than the increase in the case of high pressure and high temperature
(Figures 4.19 and 4.23 for samples one and two, respectively). This
indicates that the vaporization/condensation process was more
effective at high temperature. Therefore, it was concluded that the best
results will be obtained if the vaporization/condensation process is

76
conducted at the high depleted pressure and high vaporization
temperature.
Variation in volume percent (saturation) of liquid and vapor with
every vaporization/condensation cycle was plotted for every depleted
pressure-vaporization temperature combination for both samples
(Figures 4.27 to 4.34). Similar trends were observed in the variation of
liquid and vapor saturation as in the cases of molar variation. In the
case of sample one, the volume percent of liquid (liquid saturation)
was increased from 21% to 90% at the high pressure and high
temperature (Figure 4.27), while at the same pressure and low
temperature the liquid saturation was increased from 21% to 59%
(Figure 4.28). In the case of sample two, the liquid saturation was
increased from 19% to 80% at high pressure and high temperature
(Figure 4.31), while at the same pressure and low temperature the
liquid saturation was changed from 19% to 49% (Figure 4.32). This
again suggests that the vaporization/condensation was more effective
at high vaporization temperature. The liquid saturation was increased
from 19% to 71% at low pressure and high temperature in case of
sample one (Figure 4.29), and from 15% to 58% in case of sample two
(Figure 4.33). This increase was significant but still less than the
increase in the case of high pressure and high temperature (Figures
4.27 and 4.31 for samples one and two, respectively). This indicates
that the vaporization/condensation process was more effective at the
higher vaporization temperature. Again, it was concluded that the
best results can be obtained if the vaporization/condensation process
is conducted at the higher depleted pressure and high vaporization
temperature. The increasing trend in liquid saturation suggests that

77
after multiple cycles, the liquid saturation may be much larger than the
residual liquid saturation for a particular media. At this stage, the
liquid may become mobile and, therefore, result in additional
recovery.
Variation in overall composition with vaporization/condensation
cycles (Figures 4.35 to 4.42) were also studied. It was found that mole
percent of the lighter components (Ci, C2) was decreasing, while that
of heavier components (SCN fractions) was increasing. The variation
in mole percent of middle components (Ca-Ce) was insignificant.
Again, these trends were more pronounced in cases of high pressure
and high vaporization temperature conditions (Figures 4.35 and 4.39)
and less pronounced at low pressure and low temperature (Figures
4.38 and 4.42).
Variation trends in liquid composition (Figures 4.43 to 4.50) and vapor
composition (Figures 4.51 to 4.58) were also studied, and it was found
that there was a perceptible decrease in mole percent of the plus
fraction in the liquid composition with cycles, while decrease in the
mole percent of methane was very marginal. The change in mole
percent of other components was found to be insignificant. This
indicated that the liquid was getting lighter with every cycle. Since
lighter liquids generally have lower viscosity compared to heavier
liquids, the mobility of the liquid will increase with cycles, thus
improving the possibility of the liquid to become mobile. In the case of
vapor composition, it was found that the mole percent of methane was
decreasing with cycles, while the mole percent of other components
was increasing. This increase was significant in the case of ethane.

78
propane and plus fractions. This indicates that the gas is getting richer
with every cycle.
Trends in variation of total number of moles in place with each cycle
(Table 4.12) were also observed. It was found that the total number of
moles were increasing with each cycle at low pressures while total
number of moles were decreasing at high pressures. This observation
was contrary to expectations that the total number of moles in place
will increase with each cycle in all cases. This expectation was made
because at the vaporization stage, a large portion of the liquid should
vaporize. Placing only the enriched gas in place with original liquid
should result in an increase in overall number of moles at the new
location. It seems that this anomaly was occurring due to assumption
that the produced gas from the vaporization stage displaces
completely the gas in place at the new location. Therefore, it is
recommended that an effort should be made to improve this
assumption.
In this study, the characterization of the vaporization/condensation
process was computer modeled as a PVT cell where it was possible to increase
the temperature of the cell by heating the cell from the outside. This type of
heating does not interfere with the overall composition of the fluid in the PVT
cell. However, in the case of actual reservoir, the temperature of the reservoir
can only be raised by either injecting the heat into the reservoir (steam injection,
hot water injection) or by creating the heat in-situ (in-situ combustion). Both of
these heating mechanisms can interfere with the composition of fluid in the
reservoir. In the case of steam injection, steam will move with produced gases at
the vaporization stage and condense into water at the condensation stage. In the
case of in-situ combustion, large amounts of combustion gases will be generated

79
at the vaporization stage along with thermal cracking of oil thus, affecting the
overall composition. Therefore, the effect of the heating mechanism should be
taken into consideration when conducting this study for an actual reservoir. The
effects of flow characteristics, primarily viscosity, of the fluids (gas and liquid) in
the reservoir should also be kept in mind when characterizing the
vaporization/condensation process for an actual reservoir.

80
Table 4.1
Composition and Properties of SCN Fractions
(Sample 1)

SCN Composition MW SG Tb Pc Tc CO

Fraction (Mole %) (K) (atm) (K)


C7-C10 4.38238 113.4 0.723 384.9 27.0 563.5 0.3601
C11-C13 2.08084 161.6 0.772 461.1 20.7 641.6 0.5153
C14-C16 1.32715 207.7 0.801 519.5 17.1 697.3 0.6461
C17-C20 0.79125 257.3 0.821 573.3 14.4 745.8 0.7726
C21+ 0.91839 371.0 0.851 676.7 10.7 833.1 1.0081

Figure 4.2
Composition and Properties of SCN Fractions
(Sample 2)

SCN Composition MW SG Tb Pc Tc CO

Fraction (Mole %) (K) (atm) (K)

C7-C9 3.84551 108.8 0.751 383.1 29.9 569.8 0.3428

C10-C12 1.40394 148.7 0.801 450.6 23.8 641.5 0.4699

C13-C14 0.80174 182.9 0.829 498.1 20.5 689.1 0.5685

Cl5+ 0.79882 257.3 0.871 583.2 16.0 769.5 0.7535

81
Table 4.3
Comparison of Retrograde Liquid Volume Calculated By
Different EOS (Sample 1)

Pressure Retrograde Liquid Volume


(psia) (% of HC Pore Volume)
Experimental Calculated
PREOS SRKEOS SRKGDEOS
5102 0.0 0 7.879 7.886
5098 0.6 0 7.882 7.888
5075 3.5 3.307 7.896 7.902
5015 9.2 11.909 7.932 7.936
4918 14.7 18.477 7.992 7.993
4723 20.9 24.1 8.118 8.113
4415 26.0 27.508 8.348 8.333
3615 30.8 29.63 9.671 9.616
2715 30.6 28.29 22.575 22.413
1815 27.7 25.464 23.979 23.791

1215 24.9 23.197 22.471 22.295

715 22.6 21.117 20.648 20.48

15 18.5 13.744 12.222 12.048

82
Table 4.4
Comparison of Retrograde Liquid Volume Calculated By
Different EOS (Sample 2)

Pressure Retrograde Liquid Volume


(psia) (% of HC Pore Volume)
Experimental Calculated
PREOS SRKEOS SRKGDEOS
4014.7 0.0 0.0 0.0 0.0
3514.7 3.3 3.335 3.32 3.324
2914.7 19.4 20.332 21.175 22.134
2114.7 23.9 24.782 25.31 25.326
1314.7 22.5 22.79 22.226 22.122
605.7 18.1 18.002 17.094 16.949
14.7 12.6 9.989 7.45 7.247

83
Table 4.5
Comparison of Produced Gas Calculated By
Different EOS (Sample 1)

Pressure Produced Gas


(psia) (% of Original Mole)
Experimental Calculated
PREOS SRKEOS SRKGDEOS
5102 0.0 0.0 19.042 19.454
4415 6.955 7.602 24.039 24.444
3615 18.709 18.979 31.467 31.854
2715 34.922 34.834 44.963 45.303
1815 53.555 53.259 60.518 60.749
1215 66.298 66.355 71.565 71.72
715 77.132 77.43 81.044 81.14

84
Table 4.6
Comparison of Produced Gas Calculated By
Different EOS (Sample 2)

Pressure Produced Gas


(psia) (% of Original Mole)
Experimental Calculated
PREOS SRKEOS SRKGDEOS
4014.7 0.0 0.0 0.0 0.0
3514.7 5.374 4.341 1.887 1.671
2914.7 15.438 12.189 11.638 11.481
2114.7 35.096 33.33 31.343 31.073
1314.7 57.695 54.681 51.681 51.356
605.7 76.787 75.804 73.346 75.116
14.7 93.515 95.488 96.262 96.196

Table 4.7
Comparison of Saturation Pressures Calculated By
Different EOS

Saturation Pressure
(psia)
Experimental Calculated
PREOS SRKEOS SRKGDEOS

Sample 1 5102.0 5089.1 9349.6 9587.4

Sample 2 4014.7 3596.6 3574.4 3559.0

85
Table 4.8
Comparison Between Experimentally Obtained
Composition and Composition Calculated By
PREOS at Depleted Reservoir Pressure
(Sample 1)

Component Composition at 715 psia, 264 F


(Mole %)
Liquid Vapor
Experimental Calculated Experimental Calculated
CO2 0.68 0.65 2.46 2.50
N2 0.07 0.07 0.77 0.76
Ci 12.73 13.45 75.83 74.80
C2 3.88 3.63 8.20 8.60
C3 3.44 3.32 3.94 4.11
i-Q 2.14 2.20 1.65 1.69

n-C4 2.99 2.85 1.91 1.85

i-Cs 2.94 2.94 1.23 1.18

n-Cs 1.97 1.91 0.73 0.68

C6 5.67 6.04 1.23 1.28

C7-H 63.49 62.93 2.05 2.55

86
Table 4.9
Comparison Between Experimentally Obtained
Composition and Composition Calculated By
PREOS at Depleted Reservoir Pressure
(Sample 2)

Component Composition at 619.7 psia, 186 F


(Mole %)
Liquid Vapor
Experimental Calculated Experimental Calculated
CO2 0.05 0.05 0.21 0.20
N2 0.01 0.01 0.14 0.14
Ci 9.87 10.42 66.59 67.28
C2 7.13 7.26 16.06 15.85
C3 9.21 9.48 9.11 9.09
i-C4 1.80 1.81 1.01 0.96
n-O 7.54 7.47 3.31 3.19

i-Cs 2.74 2.69 0.68 0.63


n-Cs 5.42 5.18 1.02 1.02

Ce 8.03 8.27 0.80 0.86

C7-H 48.20 47.37 1.07 0.80

87
Table 4.10
Calculated Compositions at Depleted Reservoir
Pressures and Original Reservoir Temperature
(Sample 1)

Component Composition
(Mole %)
at 750 psia at 250 psia
Overall Liquid Vapor Overall Liquid Vapor
CO2 1.69 0.64 2.50 0.99 0.22 2.38
N2 0.46 0.07 0.76 0.24 0.02 0.65
Ci 47.82 13.46 74.80 26.98 4.24 67.78
C2 6.41 3.63 8.59 4.14 1.46 8.94
C3 3.77 3.32 4.12 2.89 1.71 5.01
i-Q 1.92 2.20 1.69 1.73 1.39 2.34

n-C4 2.29 2.85 1.85 2.19 1.91 2.69

i-Cs 1.95 2.94 1.18 2.23 2.40 1.93

n-Cs 1.22 1.91 0.68 1.45 1.63 1.14

Ce 3.37 6.04 1.28 4.70 6.00 2.35

C7-C10 10.35 20.58 2.32 16.59 23.36 4.46

C11-C13 7.18 16.13 0.16 13.65 21.12 0.25

C14-C16 4.76 10.75 0.06 9.11 14.14 0.08

C17-C20 3.03 6.88 0.007 5.83 9.08 0.007

C21+ 3.77 8.58 0.00003 7.27 11.33 0.00001

88
Table 4.11
Calculated Compositions at Depleted Reservoir
Pressures and Original Reservoir Temperature
(Sample 2)

Component Composition
(Mole %)
at 750 psia at 250 psia
Overall Liquid Vapor Overall Liquid Vapor
CO2 0.14 0.06 0.20 0.08 0.02 0.19
N2 0.08 0.01 0.14 0.04 0.003 0.11
Ci 43.71 12.82 68.56 23.93 3.72 59.72
C2 12.29 8.23 15.55 8.20 3.46 16.59
C3 9.25 10.04 8.62 8.02 6.04 11.53
i-a 1.31 1.83 0.89 1.42 1.44 1.40

n-C4 4.96 7.46 2.95 5.82 6.35 4.88

i-Cs 1.48 2.61 0.58 2.12 2.72 1.06

n-Cs 2.75 4.99 0.94 4.10 5.43 1.75

C6 3.93 7.82 0.80 6.70 9.63 1.52

C7-C9 10.54 22.82 0.64 20.38 31.28 1.08

C10-C12 4.19 9.27 0.11 8.34 12.96 0.15

C13-C14 2.59 5.80 0.01 5.23 8.18 0.013

Cis+ 2.77 6.22 0.0005 5.60 8.77 0.0003

89
Table 4.12
Moles of Fluid in Place After Each Cycle

Cycle Moles of Fluid in Place


Sample 1 Sample 2
750 psi 750 psi 250 psi 250 psi 750 psi 750 psi 250 psi 250 psi
700 F 500 F 700 F 500 F 500 F 350 F 500 F 350 F
0 100 100 100 100 100 100 100 100
1 105.1 104.9 141.5 120.7 106.0 104.4 144.5 118.7
2 96.77 101.7 170.5 131.6 102.1 100.1 175.7 127.2
3 84.64 95.2 190.7 138.8 94.3 94.57 200.2 133.1
4 74.31 88.67 205.6 144.0 86.89 89.28 219.7 -

5 67.315 82.84 216.7 147.8 81.21 84.63 236.5 -

6 63.23 77.97 224.9 - 77.31 80.75 248.4 -

7 61.1 74.02 230.2 - 74.82 77.61 258.9 -

8 60.06 70.90 233.6 - 73.29 75.1 267.5 -

9 - 68.47 235.6 - 72.38 73.12 274.6 -

10 - 66.6 236.6 - - - 280.5 -

11 - - 237.1 - - - 285.3 -

90
"^ "in -
O

Final ROV
"^ nn
O Exp. ROV

.. o en .
C
Relative Volume

1 f^n
1 .ou -

1 .UU -

0.50 -
1000 2000 3000 4000 5000 6000 7000 8000
Pressure (psia)

Figure 4.1
Comparison of Relative Volume Data
(Sample 1, PREOS)

91
4.00

3.50
ol

3.00
o\
Final ROV

o Exp. ROV

a, 2.50 o\
E
O
>
_>
JS <> \
^ 2.00

o \

1.50 O X

o
o
()
1.00
^t>0O(
> O o

0.50
2000 4000 6000 8000 10000
Pressure (psia)

Figure 4.2
Comparison of Relative Volume Data
(Sample 1, SRKEOS)

92
4.00 1

3.50
ol

3.00
o O Exp. ROV

0) 2.50
E
O
>
0)
>
0) () \
C. 2.00

o \
1.50

O
o
(

1.00 ^*JOo <


' o o

0.50
2000 4000 6000 8000 10000
Pressure (psia)

Figure 4.3
Comparison of Relative Volume Data
(Sample 1, SRKGDEOS)

93
<)
3.50

"^ nn . Final ROV

O Exp. ROV

0) 2.50
E
3
O
>
Relative
3

1.50 \o

1.00 ^ ^22t^o_o_oj L^
-^-J )

0 50 -
1000 2000 3000 4000 5000 6000 7000
Pressure (psia)

Figure 4.4
Comparison of Relative Volume Data
(Sample 2, PREOS)

94
4.00

1000 2000 3000 4000 5000 6000 7000

Pressure (psia)

Figure 4.5
Comparison of Relative Volume Data
(Sample 2, SRKEOS)

95
4.00

)
3.50

3.00

I 2.50 nal ROV


_3
O
> O E Kp. ROV

SS \o
0)
a. 2.00

1.50 V o

\sO
\v4
1.00
2222o__o__oj
^.___o_j 5

0.50 4
1000 2000 3000 4000 5000 6000 7000
Pressure (psia)

Figure 4.6
Comparison of Relative Volume Data
(Sample 2, SRKGDEOS)

96
1.10

o
1
1.05

o 1.00
n> O Exp. Gas Z
u.

(/)
V)
a.
E
o
o
\
0) 0.95
(0
O o
/

o
O /
0.90
o /

'

0.85
1000 2000 3000 4000 5000 6000
Pressure (psia)

Figure 4.7
Comparison of Z-Factor Data
(Sample 1, PREOS)

97
1.40
1

1.30

Z
1.20
O Exp. Gas Z
u

1.10 /
<1>
Q.
E 1 /
o
o / o
(A
(Q
i /
o
1.00 I

\^^ o
o 1
o

0.90 r>
o

0.80
1000 2000 3000 4000 5000 6000
Pressure (psia)

Figure 4.8
Comparison of Z-Factor data
(Sample 1, SRKEOS)

98
1.40

1.30
1

\ /

Final GasZ
1.20 f
o Exp. Gas Z

u
(Q

/
1

1.10
a ! ' / '
E : / 1
o
o
(A
ra
0

1.00

^^^ o
o^ o

0.90 o
o

0.80
1000 2000 3000 4000 5000 6000
Pressure (psia)

Figure 4.9
Comparison of Z-Factor Data
(Sample 1, SRKGDEOS)

99
1.00

!
0.95
i

1
O Exp. Gas Z

0.90 \ o
1
u \
ra
u.
0
1

J8 0.85

a.
E
o
o
(A
ra
O \
0.80

/
\ O /
/
/
0.75
^^ or
1

i
i

0.70
1000 2000 3000 4000 5000
Pressure (psia)

Figure 4.10
Comparison of Z-Factor Data
(Sample 2, PREOS)

100
1.00

0.95

-rmeai \jci9 t-

o Exp. Gas Z

0.90

u /
ra \
u.
/
i;; 0.85
0) i
a 1
E
o I
o 1
(A
ra o \
\ 1

0.80
/ o
j ^ ^ ^

1
1 o
1
0.75 1
o

i
0.70
1000 2000 3000 4000 5000
Pressure (psia)

Figure 4.11
Comparison of Z-Factor Data
(Sample 2, SRKEOS)

101
1.00

1
0.95 -
1
i

O Exp. Gas Z

0.90
\ i
u
ra
u. \ 1
\
i \
1 o
21 0.85 i \

Q.
E
o
O
(A
ra o \
CD
0.80
u

0
0.75
o

0.70
1000 2000 3000 4000 5000
Pressure (psia)

Figure 4.12
Comparison of Z-Factor Data
(Sample 2, SRKGDEOS)

102
35.0 100.0

\
\
\ A
\ A 1
30.0 1
\ 1 .^0'^'^ ^^*N.
\
\ 80.0
\ A
A Exp. Liq. Vol. i
\ A\
\ Final Prod. Gas
25.0 A /^
\ /
O Exp. Prod. Gas
A
\
o \ 1
I I
> 60.0 a
"ra
E 20.0
\
\ A\ 1
1 1
c \ a
o A / O
(A

E \
_3 \
O (Q
> 15.0 \ 3
D \ L 0)
\ 1 40.0 -
V 1 o

\
\
10.0 V
\ L^
\
< \
\
20.0
\
5.0 V
\
\

b.

0.0 0.0
1000 2000 3000 4000 5000 6000 7000
Pressure (psia)

Figure 4.13
Comparison of CVD Test Data
(Sample 1, PREOS)

103
40.0 100.0

\
35.0
\
\ Final Liq. Vol. 80.0
s
O^
30.0 A Exp. Liq. Vol.

Final Prod. Gas


o ^
o 25.0 ^ ^ O Exp. Prod. Gas
>
c
"5)
'LI
O

20.0
0)
E
_3
O
>
o
"3
tr

1000 2000 3000 4000 5000 6000 7000


Pressure (psia)

Figure 4.14
Comparison of CVD Test Data
(Sample 1, SRKEOS)

104
40.0 100.0
1

35.0
Final Liq. Vol. |
\
* i
^\ j A Exp. Liq. Vol.
80.0
1 i
o *\ A A Final Prod. Gas
30.0 1

O Exp. Prod. Gas


A
0 '\
% A
o 25.0 A
A
> k 60.0 I
"ra y ^
\^ o
c A (D
"5) ^ o A
a
O
20.0 fi)
0) (0

E
_3
A f X
1
1
1
X
O X

> >
"5 40.0^
2" 15.0 3
\x A o
% X
O \ X

\ \
\
\\
\ s
\ s
\
10.0

20.0
o

5.0
A
O

0.0 0.0
1000 2000 3000 4000 5000 6000
Pressure (psia)

Figure 4.15
Comparison of CVD Test Data
(Sample 1, SRKGDEOS)

105
30.0 1
100.0

1
' ^*s,^^^

25.0 X
Final Liq. Vol. ^
X

\ A \ 80.0
\ A Exp. Liq. Vol.
'9
1

\ Final Prod. Gas


.
20.0
^ \\ A\ O Exp. Prod. Gas
X
o X
> i \ 60.0 .
75 \ O I\ o
1 \ (D
c 1 1 a
'5) \ O
u
o 15.0 \ ! \ (A
0) V

E X
1 1

_3 X
X
! I
O
> A/
X
i \
'5
X
X i 1
40.0 ^
cr X 1 \ 3
o
o
10.0 %
\
X

\
X
X
X
\
X
X
X

\ 20.0
5.0
\0 \
%
X
X X \

1 ^ ^v
1 ^ ^ .
0.0 - " ^ ^j 0.0
1000 2000 3000 4000 5000
Pressure (psia)

Figure 4.16
Comparison of CVD Test Data
(Sample 2, PREOS)

106
30.0 100.0

25.0 -

20.0
o
>
75
c
'5
0)
15.0
E
o
>
"5

10.0

1000 2000 3000


Pressure (psia)

Figure 4.17
Comparison of CVD Test Data
(Sample 2, SRKEOS)

107
30.0 100.0
t

>
1

\ i

25.0
X
X
A N. riRal Liq. vol. 80.0
\
\
\ 1
p
V
\
X
A Exp. Liq. Vol.
\

X

20.0 * Final Prod. Gas


X

k
A
o /
> A / 0 Exp. Prod. Gas 60.0 Q.
c
75 \ 0 o
c
X a
O
"5) \\ 0)
1
V)
15.0 \
E
X \
_3
O
X
X
\ CO
> A / X
3'
"3 \ 40.0 -
X
X o
X
0
10.0
\
X
X
X
X

X
X
X

X I
20.0
X 1
X 1
5.0
1 -.O; \
! ! \
A
X \

| \ o\
X ^V
\ ^k

0.0 ! ' ^^ , 0.0


1000 2000 3000 4000 5000
Pressure (psia)

Figure 4.18
Comparison of CVD Test data
(Sample 2, SRKGDEOS)

108
100
4 f 1

^^,0^''''^ T

90

80

70

60 i

"""Mole % of Liquid
I 50
- * - M o l e % of Vapor 1

\
40

30

20

10

J "11

4 8
Cycles

Figure 4.19
Variation in Mole Percent of Liquid and Vapor
(Sample 1, 750 psia, 700 F)

109
90

80

70

60

Mole % of Liquid
50
Mole % of Vapor
0)
o
40

30

20

10

4 6 8 10
Cycles

Figure 4.20
Variation in Mole Percent of Liquid and Vapor
(Sample 1, 750 psia, 500 F)

110
100
- .,

90

80

70

iV

60

""Mole % of Liquid
I 50
n Mole % of Vapor

40

30

20
:

10 1

6 8 10 12
Cycles

Figure 4.21
Variation in Mole Percent of Liquid and Vapor
(Sample 1, 250 psia, 700 F)

111
20

10

r-
4
Cycles

Figure 4.22
Variation in Mole Percent of Liquid and Vapor
(Samplel, 250 psia, 500 F)

112
100 1

jO)
o

4 5 8
Cycles

Figure 4.23
Variation in Mole Percent of Liquid and Vapor
(Sample 2, 750 psia, 500 F)

113
80

' ^^^-^ )y

70

i
60

Mole % of Liquid
50
n Mole % of Vapor

I 40

^^"^^-"--^
30 ^ i

20

10

4 5 8
Cycles

Figure 4.24
Variation in Mole Percent of Liquid and Vapor
(Sample 2, 750 psia, 350 F)

114
100

90
!

80

70

f
60

6^

I 50 Mole % of Liquid

OMole % of Vapor
40

30

20

10

0
0 6 8 10 12
Cycles

Figure 4.25
Variation in Mole percent of Liquid and Vapor
(Sample 2, 250 psia, 500 F)

115
80

70

60

Mole % of Liquid
50
'Mole % of Vapor

I 40

30 -Q
-^

20

10

0.5 1.5 2.5

Cycles

Figure 4.26
Variation in Mole Percent of Liquid and Vapor
(Sample2, 250 psia, 350 F)

116
90

i
1
80

70

60

""Volume % of Liquid
50
0) BVolume % of Vapor
E
O
>
40

30
i
t

i
20 j

1
i

10

4 8
Cycles

Figure 4.27
Variation in Volume Percent of Liquid and Vapor
(Sample 1, 750 psia, 700 F)

117
80

70

60

^ ^ ^ " " ^

50

^^"^"-n
0)
I 40
"o
>

30

Volume % of Liquid
20
*Volume % of Vapor

10

8 10
Cycles

Figure 4.28
Variation in Volume Percent of Liquid and Vapor
(Sample 1, 750 psia, 500 F)

118
90

80

70

60

Volume % of Liquid
50
0) fiVolume % of Vapor
E
O
>
40

30

20

10

6 8 10 12
Cycles

Figure 4.29
Variation in Volume Percent of Liquid and Vapor
(Sample 1, 250 psia, 700 F)

119
0)
E
O
>

Cycles

Figure 4.30
Variation in Volume Percent of Liquid and Vapor
(Sample 1, 250 psia, 500 F)

120
90

80

70

60

Volume % of Liquid
50
0) ""BVolume % of Vapor ;
E
O
> 40

30

M
20

10

4 5 8
Cycles

Figure 4.31
Variation in Volume Percent of Liquid and Vapor
(Sample 2, 750 psia, 500 F)

121
90
:

80
'

70

60

Ji^,,^^^^

50
6^
0)
E
O
> 40

Volume % of Liquid
30
-*Volume % of Vapor

20

10

4 5 8
Cycles

Figure 4.32
Variation in Volume Percent of Liquid and Vapor
(Sample 2, 750 psia, 350 F)

122
0)
E
O
>

6 8 10 12
Cycles

Figure 4.33
Variation in Volume Percent of Liquid and Vapor
(Sample 2, 250 psia, 500 F)

123
90

i
1

ia u
80

i
70

60

Volume % of Liquid
50
0) OVolume % of Vapor
E
O
> 40

30

20 1 - <i **

10

0.5 1.5 2.5

Cycles

Figure 4.34
Variation in Volume Percent of Liquid and Vapor
(Sample 2, 250 psia, 350 F)

124
Figure 4.35
Variation in Overall Composition
(Sample 1, 750 psia, 700 F)

125
6 8 10 12
Cycles

Figure 4.36
Variation in Overall Composition
(Sample 1, 750 psia, 500 F)

126
30

14

Figure 4.37
Variation in Overall Composition
(Sample 1, 250 psia, 700 F)

127
30 T

2 3 4
Cycles

Figure 4.38
Variation in Overall Composition
(Sample 1, 250 psia, 500 F)

128
Ji
o

8 10
Cycles

Figure 4.39
Variation in Overall Composition
(Sample 2, 750 psia, 500 F)

129
45

40

35

30

25

0)
o

20

8 10
Cycles

Figure 4.40
Variation in Overall Composition
(Sample 2, 750 psia, 350 F)

130
J)
o

Figure 4.41
Variation in Overall Composition
(Sample 2, 250 psia, 500 F)

131
J)
o

Cycles

Figure 4.42
Variation in Overall Composition
(Sample 2, 250 psia, 350 F)

132
70 1

002

N2
50
01
02
03
40
104

o N04

105
30
N05

F06

07+
20

Figure 4.43
Variation in Liquid Composition
(Sample 1, 750 psia, 700 F)

133
70

60

002
50
N2

01

02
40 -
03

o 104

N04
30
105

N05

20 F06

07+

10

Figure 4.44
Variation in Liquid Composition
(Sample 1, 750 psia, 500 F)

134
an

70

002
60
N2

01
50
02

03

I 40 104

N04

105
30
N05

F06
20
07+

10

12 14

Figure 4.45
Variation in Liquid Composition
(Sample 1, 250 psia, 700 F)

135
3
Cycles

Figure 4.46
Variation in Liquid Composition
(Samplel, 250 psia, 500 F)

136
002

35 N2

01

02
30
03

104
25
N04
J)
o 105
20
N05

F06

15 07+

8 10

Cycles

Figure 4.47
Variation in Liquid composition
(Sample 2, 750 psia, 500 F)

137
o

8 10
Cycles

Figure 4.48
Variation in Liquid Composition
(Sample 2, 750 psia, 350 F)

138
70 1

50 002

N2

01

40 02

03
o
104
30
N04

105

N05
20
F06

07+

10

0 8 10
Cycles

Figure 4.49
Variation in Liquid Composition
(Sample 2, 250 psia, 500 F)

139
70

60

002
50
N2

01

40 02

03
J)
o
104
30 N04

105

N05
20
F06

07+

1 2
Cycles

Figure 4.50
Variation in Liquid Composition
(Sample 2, 250 psia, 350 F)

140
80 T

70

60

50

I 40

30

20

Figure 4.51
Variation in Vapor Composition
(Sample 1, 750 psia, 700 F)

141
002

60 N2

01
02
50 03
104
N04
^ 40
105

N05

30 F06

07+

20

Figure 4.52
Variation in Vapor Composition
(Sample 1, 750 psia, 500 F)

142
60

50

40

J)
o

30

20

10 ^C )( K )C )( X K )( ^^ ^
:iK

12 14

Figure 4.53
Variation in Vapor Composition
(Sample 1, 250 psia, 700 F)

143
70 1

002

50 N2

01

02
40
03

o 104

N04
30
105

N05

F06

07+

3
Cycles

Figure 4.54
Variation in Vapor Composition
(Samplel, 250 psia, 500 F)

144
70

60 002

N2

01
50 02

03

104
40
N04
J) 105
o
N05
30
F06

07+

20

10 .)K 3 F ^ )K rt H Nt
:i *-

8 10
Cycles

Figure 4.55
Variation in Vapor Composition
(Sample 2, 750 psia, 500 F)

145
70

60
0O2

N2

50 01

02

03

40 104

N04
o 105

30 N05

F06

07+

20

^( X ?( X X -X

0
8 10
Cycles

Figure 4.56
Variation in Vapor Composition
(Sample 2, 750 psia, 350 F)

146
50 002

N2

01

02
40
03

104

N04

I 30 105

N05

F06

20 07+

10

0
4 6 8 10

Cycles

Figure 4.57
Variation in Vapor Composition
(Sample 2, 250 psia, 500 F)

147
60

50

002

N2
40 01

02

03

i 30 104

N04

105

N05
20
F06
^

07+

10

1 2
Cycles

Figure 4.58
Variation in Vapor Composition
(Sample 2, 250 psia, 350 F)

148
CHAPTER 5
CONCLUSIONS

Based on this study, the following conclusions have been drawn:


1. The vaporization/condensation process was found to be feasible.
2. The Peng-Robinson equation of state model was identified as the
most accurate model to simulate experimental data.
3. The vaporization/condensation process resulted in increased liquid
saturation. If this increased saturation was more than the residual
liquid saturation in terms of relative permeability, then liquid may
become mobile, resulting in additional recovery.
4. The vaporization/condensation process resulted in increased mole
percent of liquid and decreased mole percent of vapor.
5. The vaporization/condensation process was found to be more
effective at high reservoir pressure and high vaporization
temperature.
6. The vaporization/condensation process resulted in the decreased
mole percent of lighter components (CI, C2) and increased mole
percent of heavier components (SCN fractions) in the overall fluid
composition. The effect on middle components (C3-C6) was
insignificant.
7. Liquid and vapor compositions were affected in such a way by the
vaporization/condensation process that liquid was lighter and
vapor was enriched after the vaporization/condensation process.

149
CHAPTER 6
RECOMMENDATIONS

Based on the present stiidy, the following are the recommendations:


1. This study was conducted on only two samples of a particular fluid
type (gas condensate). Fuhire studies should be conducted on
large number of samples of different type of fluid (black oil, volatile
oil etc.) to make the study more general and the study results more
relevant.
2. Only three equation of state models were tested in this study.
Other more recent equation of state models, such as Patel-Teja,
Schmidt-Wenzel equation of state model should be tested to verify
the results.
3. It was assumed in this study that the temperature increases
suddenly from initial temperature to the specified vaporization
temperature at the vaporization stage and decreases suddenly from
vaporization temperature to initial temperature at the condensation
stage. In the actual reservoir, however, any increase or decrease of
temperature happens gradually. Therefore, an effort should be
made to quantify this assumption.
4. This study was conducted in context of a PVT cell. When
conducting the study for an actual reservoir, the flow
characteristics of the fluids in the reservoir and effects of heating
mechanism (steam injection, in-situ combustion) should be
considered.

150
REFERENCES

1. Abbott, M. M., ''Cubic Equation of State: An Interpretive Review,"


American Chem. Society, Advances in Chemistry Series, Volume 182,
1979, pp. 47-70.

2. Martin, J.J., "Cubic Equation of State - Which?," Ind. Ene. Chem. Fund.,
18(2), 1979, pp. 81-97.

3. Firoozabadi, A., "Reservoir-Fluid Phase Behaviour and Volumetric


Predictions with Equation of state," Journal of Petroleum Technology,
40(4), 1988, pp. 387-406.

4. Katz, D. L., and Firoozabadi, A., "Predicting Phase Behavior of


Condensate/Crude-Oil Systems Using Methane Interaction Coefficients,"
Journal of Petroleum Technology, Volume 30, November 1978, pp. 1649-
1655.

5. Sarkar, R., Danesh, A. S., Todd, A. C , "Phase Behavior Modeling of Gas-


Condensate Fluids Using an Equation of State," SPE 22714, Presented at
the 66*^ Annual Technical Conference of the Society of Petroleum
Engineers held in Dallas, Texas, October 6-9,1991.

6. Ahmed, T. H., "Comparative Study of Eight Equations of State for


Predicting Hydrocarbon Volumetric Phase Behavior," SPE 15673,
Presented at the 61* Annual Technical Conference of the Society of
Petroleum Engineers held in New Orleans, LA, October 5-8,1986.

7. Fassihi, M. R., "Improved Phase Behavior Representation of Thermal


Recovery of Light Oils," SPE 24034, Presented at Western Regional
Meeting of Society of Petroleum Engineers held in Bakersfield, California,
March 30-April 1,1992.

8. Danesh, A., Dong-Hai, Xu, and Todd, A. C , " An Evaluation of Cubic


Equations of State for Phase Behavior Calculations Near Miscibility
Conditions," SPE/DOE 20267, Presented at SPE/DOE Seventh
Symposium on Enhanced Oil Recovery held in Tulsa, Oklahoma, April 22-
25,1990.

151
9. Bjorlykke, O. P., and Firoozabadi, A., "Measurement and Computation of
Retirograde Condensation and Near-Critical Phase Behavior," SPE 20524,
Presented at the 65* Annual Technical Conference of the Society of
Petroleum Engineers held in New Orleans, LA, September 23-26,1990.

10. Frailey, Scott M., Personal Communication. Assistant Professor in the


Department of Petroleum Engineering. Texas Tech University, Lubbock,
Texas, 1996-97.

11. Amyx, James W., Bass, Daniel M., Jr. and Whiting, Robert L., Petroleum
Reservoir Engineering - Physical Properties, McGraw-Hill Classic
Textbook Reissue, McGraw-Hill Book Company, New York, 1988

12. Campbell, John M., Gas Conditioning and Processing, Volume 1: The
Basic Principles, Seventh Edition, Campbell Petroleum Series, Norman,
Oklahoma, 1992.

13. McCain, W. D., Jr., The Properties of Petroleum Fluids, Volume 2,


PennWell Publishing Company, Tulsa, Oklahoma, 1990.

14. Ahmed, T., Hydrocarbon Phase Behavior, Volume 7, Gulf Publishing


Company, Houston, Texas, 1989.

15. Wilson, G., "A Modified Redlich-Kwong EOS, Application to General


Physical Data Calculations," Paper 15C, Presented at the Annual AIChE
National Meeting held in Cleveland, Ohio, May 4-7,1968.

16. Standing, M. B., "A set of Equations for Computing Equilibrium Ratios of
a Crude Oil/Natural Gas System Below 1000 psia," Journal of Petroleum
Technologv, September 1979, pp. 1193-1195.

17. Leland, T. W., "Phase Equilibria and Fluid Properties in the Chemical
Industry," DECHEMA, 1980, pp. 283-333.

18. Benedict, W., Webb, G. B., and Rubin, L. C , "An Empirical Equation for
Thermodynamic Properties of Light Hydrocarbons and Their Mixtures. I.
Methane, Ethane, Propane, and Butane/' J. Chem. Phys., Volume 8,1940,
pp. 334-345.

19. Sterling. K.. Fluid Thermodvnamic Properties for Light Petroleum


Systems, Gulf Publishing Company, Houston, Texas, 1973.

152
20. Lee, B. I., and Kesler, M. G., "A Generalized Thermodynamic Correlation
Based on Three-Parameter Corresponding States," AIChE Journal,
Volume 21,1975, pp. 510-527.

21. Tsonopoulos, C , and Heidman, J. L., "High Pressure Vapor-Liquid


Equilibria with Cubic Equations of State," Fluid Phase Equilibria, Volume
29,1986, pp. 391-414.

22. Peng, D., and Robinson, D., "A New Two Constant Equation of State,"
Industrial and Engineering Chem. Fund., 1976, Vol. 15, No. 1, pp. 59-64.

23. Redlich, O., and Kwong, J., "On The Thermodynamics of Solutions. An
Equation of State . Fugacities of Gaseous Solutions," Chemical Reviews,
Vol. 44,1949, pp. 233-247.

24. Zudkevitch, D., and Joffe, J., "Correction and Prediction of Vapor-Liquid
Equilibria with the Redlich-Kwong Equation of state," AIChE Tournal,
Volume 16(1), 1970.

25. Joffe, J., Schroeder, G. M., and Zudkevitch, D., "Vapour-Liquid


Equilibrium with the Redlich-Kwong Equation of State," AIChE Journal,
1970, pp. 496-98.

26. Soave, G., "Equilibrium Constants from a Modified Redlich-Kwong


Equation of State/' Chem. Eng. Sci., 1972, Vol. 27, pp. 1197-1203.

27. Edmister, W., and Lee, B., Applied Hydrocarbon Thermodvnamics,


Volume 1, Second Edition, Gulf Publishing Company, Houston, Texas,
1986.

28. Slot-Petersen, C , "A Systematic and Consistent Approach to Determine


Binary Interaction Coefficients for the Peng-Robinson Equation of State,"
Paper SPE 16941, Presented at the 62"^^ Annual Technical Conference of
the SPE, held in Dallas, Texas, September 27-30,1987.

29. Vidal, J., and Daubert, T., "Equation of State - Reworking the Old Forms,"
Chem. Eng. Sci., 1978, Vol. 33, pp. 787-791.

30. Graboski, M. S., and Daubert, T., "A Modified Soave Equation of State for
Phase Equilibrium Calculations 1. Hydrocarbon System," Ind. Eng. Chem.
Process Pes. Dev., 1978, Vol. 17, pp. 443-448.

153
31. Elliot, J., and Daubert, T., "Revised Procedure for phase Equilibrium
Calculations with Soave Equation of State," Ind. Eng. Chem. Process Pes.
Dev., 1985, Volume 23, pp. 743-748.

32. Peneloux, A., Rauzy, E., and Freze, R., "A Consistent Correlation for
Redlich-Kwong -Soave Volumes," Fluid Phase Equilibria, 1982, Vol. 8, pp.
7-23.

33. Nikos, v., et al., "Phase Behavior of Systems Comprising North Sea
Reservoir Fluids and Injection Gases," Journal of Petroleum Technology,
November 1986, pp. 1221-1233.

34. Stryjek, R., and Vera, J. H., "PRSV: An Improved Peng-Robinson Equation
of State for Pure Compounds and Mixtures," Canadian J. Chem. Eng.,
April 1986, Volume 64, pp. 323-333.

35. Stryjek, R., and Vera, J. H., "PRSV: An Improved Peng-Robinson Equation
of State with New Mixing Rules for Strongly Nonideal Mixtures,"
Canadian J. Chem. Eng., April 1986, Volume 64, pp. 334-340.

36. Jhaveri, B. S., and Youngren, G. K., "Three-Parameter Modification of the


Peng-Robinson Equation of State to Improve Volumetric Predictions,"
Paper SPE 13118, Presented at the 1984 SPE Annual Technical Conference,
Houston, Texas, September 16-19.

37. Ahmed, T., "A Practical Modification of the Peng-Robinson Equation of


State," Paper SPE 18532, Presented at the SPE Eastern Regional Meeting,
Charleston, West Virginia, November 1-4,1988.

38. Schmidt, G., and Wenzel, H., "A Modified Van der Waals Type Equation
of State," Chem. Eng. Sci., 1980, Volume 135, pp. 1503-1512.

39. Patel, N., and Teja, A., "A New Equation of State for Fluids and Fluid
Mixtures," Chem. Eng. Sci. Volume 37, No. 3, pp. 463-473,1982.

40. Willman, B. T., and Teja, A. S., "Continuous Thermodynamics of Phase


Equilibria Using a Multivariable Distribution Function and An Equation
of State," AIChE Tournal, December 1986, Volume 32, No. 12, pp. 2067-
2078.

41. Valderrama, J. O., and Cisternas, L., "A Cubic Equation of State for Polar
and Other Complex Mixtures," Fluid Phase Equilibria, Volume 29, pp.
431-438,1986.

154
42. Valderrama, J. O., "A Generalized Patel-Teja Equation of State for Polar
and Non-Polar Fluids and Their Mixhires," J. Chem. Eng., Volume 23(1),
1990, pp. 87-91.

43. Pedersen, K. S., Fredenslund, Aa., and Thomassen, P., Properties of Oils
and Natural Gases. Gulf Publishing Company, Houston, Texas, 1989.

44. Reid, Robert C , Prausnitz, John M., Poling, Bruce E., The Properties of
Gases & Liquids, Fourth Edition, McGraw-Hill Book Company, New
York, USA, 1987.

45. Danesh, A., Dong-Hai, Xu, and Todd, A. C , "Comparative Study Of Cubic
Equation of State for Predicting Phase Behavior and Volumetric Properties
of Injection Gas-reservoir Oil Systems," Fluid Phase Equilibria, Volume
63,1991, pp. 259-278.

46. Coats, K. H., and Smart, G. T., "Application of a Regression Based EOS
PVT Program to Laboratory Data," SPE Reservoir Engineering, May 1984,
pp. 277-299.

47. Agarwal, R. K., Li, Y., and Nghiem, L., "A Regression Technique with
Dynamic Parameter Selection for Phase Behavior Matching," SPE
Reservoir Engineering, February 1990, pp. 115-120.

48. Whitson, C.H., "Characterizing Hydrocarbon Plus Fractions," SPE


Journal, Volume 23,1983, pp. 683-684.

49. Merril, R. C , Hartman, K. J., and Creek, J. L., "A Comparison of Equation
of State Tuning Methods," SPE 28589, Presented at the 69* Annual
Technical Conference of the Society of Petroleum Engineers held in New
Orleans, Louisiana, September 25-28,1994.

50. Merril, R. C , and Newley, T. M. J., "A Systematic Investigation into the
Most Suitable Data for the Development of Equation of State for
Petroleum Reservoir fluids," Fluid Phase Equilibria, Volume 82,1993, pp.
101-110.

51. Pedersen, K. S., Thomassen, P., and Fredenslund, Aa., "Characterization


of Gas condensate mixtures," Presented at AIChE National Meeting at
New Orleans, Louisiana, March 6-10,1988.

155
52. Pedersen, K. S., Thomassen, P., and Fredenslund, Aa., "On the Dangers of
Tuning Equations of State Parameters," Chemical Engineering Science,
Volume 43,1988, pp. 269-278.

53. Hariu, O., and Sage, R., "Crude Split Figured by Computer,"
Hvdrocarbon Proces., April 1969, pp. 143-148.

54. Katz, D., "Overview of Phase Behavior of Oil and Gas Production,"
Tournal of Petroleum Technology, June 1983, pp. 1205-1214.

55. Ahmed, T., Cady, G., and Story, A., "A Generalized Correlation for
Characterizing the Hydrocarbon Heavy Fractions," Paper SPE 14266,
Presented at the 60* Annual Technical Conference of the Society of
Petroleum Engineers held in Las Vegas, September 22-25,1985.

56. HYSIM User's Guide, Hyprotech Limited, Calgary, Alberta, Canada, 1991.

57. WinProp Phase Property Program for Windows-User's Guide, Computer


Modelling Group, Calgary, Canada, 1997.

156
PERMISSION TO COPY

In presenting this thesis in partial fulfillment of the requirements for a


master's degree at Texas Tech University or Texas Tech University Health Sciences
Center, I agree that the Library and my major department shall make it freely
available for research purposes. Permission to copy this thesis for scholarly
purposes may be granted by the Director of the Library or my major professor.
It is understood that any copying or publication of this thesis for financial gain
shall not be allowed without my further written permission and that any user
may be liable for copyright infringement.

Agree (Permission is granted.)

Stiiclefff's Signature Date

Disagree (Permission is not granted.)

Student's Signature Date

Potrebbero piacerti anche