field performance analysis and optimization of gas

202
The Pennsylvania State University The Graduate School Department of Energy and Mineral Engineering FIELD PERFORMANCE ANALYSIS AND OPTIMIZATION OF GAS CONDENSATE SYSTEMS USING ZERO-DIMENSIONAL RESERVOIR MODELS A Thesis in Energy and Mineral Engineering by Pichit Vardcharragosad 2011 Pichit Vardcharragosad Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science August 2011

Upload: others

Post on 25-Oct-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

Department of Energy and Mineral Engineering

FIELD PERFORMANCE ANALYSIS AND OPTIMIZATION OF

GAS CONDENSATE SYSTEMS USING ZERO-DIMENSIONAL RESERVOIR MODELS

A Thesis in

Energy and Mineral Engineering

by

Pichit Vardcharragosad

2011 Pichit Vardcharragosad

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science

August 2011

The thesis of Pichit Vardcharragosad was reviewed and approved* by the following:

Luis F. Ayala

Associate Professor of Petroleum and Natural Gas Engineering

Thesis Advisor

R. Larry Grayson

Professor of Energy and Mineral Engineering

Graduate Program Officer of Energy and Mineral Engineering

Li Li

Assistant Professor of Energy and Mineral Engineering

Yaw D. Yeboah

Professor of Energy and Mineral Engineering

Head of the Department of Energy and Mineral Engineering

*Signatures are on file in the Graduate School

iii

ABSTRACT

Field performance prediction is a crucial piece of information that all relevant parties

have to use in their design and decision processes during the development and exploitation of a

hydrocarbon reservoir. Field performance analysis is an engineering task which requires

knowledge, time, and the right tools and models. Available tools such as commercial reservoir

simulators might not always be the most efficient or the most optimum solution even if they make

use of highly sophisticated or detailed models. This is because these sophisticated models often

require more input data and longer running time than the less detailed ones. These problems

become worse when availability of input data and working time are constrained.

This thesis aims to develop a field performance model which will allow engineer to

perform analysis and optimization tasks more effectively for the case of gas condensate

reservoirs. A gas condensate is one of the many fluid types that can be found in conventional

hydrocarbon reservoirs. The development of a two-phase condition below the dew point pressure

can significantly increases the complexity in engineering performance calculation. The proposed

tool utilizes both zero-dimensional reservoir model customized for gas condensate and pseudo

component model. Results indicate that both models can provide fairly good prediction results

while requiring much less input and running time. Microsoft Excel with built-in Visual Basic for

Applications (VBA) is selected as the platform to develop this simulator due to the user-friendly

interface, useful built-in features, and high flexibility to use and hard-code modification. The

proposed model is able to successfully predict field performance while capturing all major fluid

behavior characteristics of gas condensates as well as being capable of performing various

optimization tasks effectively. Limitations of the implemented pseudo component model, such as

negative solution gas-oil ratios at low reservoir pressure, are elaborated and discussed. The

possible sources of error and associated preventive measures derived from the use of a gas

iv

condensate tank model for the case volatile oil reservoirs are addressed. Further recommended

studies on negative value of decline exponent variable and expanding current capability of the

proposed model are also presented.

v

TABLE OF CONTENTS

List of Figures .......................................................................................................................... vii List of Tables ........................................................................................................................... ix Nomenclature ........................................................................................................................... x Acknowledgements .................................................................................................................. xv

Chapter 1 Introduction ............................................................................................................. 1

Chapter 2 Background ............................................................................................................. 3

2.1 Gas Condensate Hydrocarbon Fluid .......................................................................... 3 2.2 Modified Black-Oil Model ......................................................................................... 5 2.3 Zero-Dimensional Reservoir Model .......................................................................... 7 2.4 Field Performance Prediction ..................................................................................... 8 2.5 Visual Basic for Applications (VBA) ........................................................................ 12

Chapter 3 Problem Statement .................................................................................................. 13

Chapter 4 Model Description ................................................................................................... 15

4.1 Phase Behavior Model (PBM) ................................................................................... 16 4.1.1 Compressibility Factor .................................................................................... 17 4.1.2 Vapor-Liquid Equilibrium ............................................................................... 21 4.1.3 Fluid Property Prediction ................................................................................ 27 4.1.4 Phase Stability Analysis .................................................................................. 37

4.2 Standard PVT Properties ............................................................................................ 42 4.2.1 Definitions, Mathematic Relationships, and Characteristics ........................... 43 4.2.2 Obtaining Standard PVT Properties from Laboratory PVT Reports ............... 51 4.2.3 Obtaining Standard PVT Properties from a Phase Behavior Model ............... 60

4.3 Zero-Dimensional Reservoir Model .......................................................................... 70 4.3.1 Generalized Material Balance Equation .......................................................... 71 4.3.2 Material Balance Equation for a Gas Condensate Fluid ................................. 75 4.3.3 Phase Saturation Calculations ......................................................................... 80 4.3.4 Volumetric OGIP/OOIP Calculations ............................................................. 82

4.4 Flow Rates and Flowing Pressures Calculation ......................................................... 84 4.4.1 Inflow Performance Relationship (IPR) .......................................................... 85 4.4.2 Tubing Performance Relationships ................................................................. 91 4.4.3 Nodal Analysis ................................................................................................ 96

4.5 Field Performance Prediction ..................................................................................... 99 4.5.1 Performance during Plateau Period ................................................................. 100 4.5.2 Performance during Decline Period ................................................................ 103 4.5.3 Annual Production Calculation ....................................................................... 106

4.6 Economic Analysis and Field Optimization............................................................... 109 4.6.1 Simplified Economic Model ........................................................................... 110 4.6.2 Field Optimization ........................................................................................... 116

vi

Chapter 5 Model Performance ................................................................................................. 117

5.1 Simulation of Standard PVT ―Black Oil‖ Properties ................................................. 117 5.1.1 Simulated Standard PVT Properties ................................................................ 117 5.1.2 Limitations of Pseudo Component Model ....................................................... 121 5.1.3 Impact on Standard PVT Properties ................................................................ 127

5.2 Zero-Dimensional Material Balance Calculations for Gas Condensates ................... 129 5.2.1 Simulation Results from Gas Condensate Tank Model .................................. 129 5.2.2 Misuse of Gas Condensate Tank Model in Volatile Oil Reservoir ................. 132

5.3 Field Performance Prediction ..................................................................................... 137 5.3.1 Field Performance Prediction Results ............................................................. 137 5.3.2 Decline Trend Analysis ................................................................................... 143

5.4 Economic Analysis and Optimization ........................................................................ 146 5.4.1 Field Economic Analysis ................................................................................ 146 5.4.2 Field Optimization ........................................................................................... 149

5.5 Application for Other Production Situations .............................................................. 152 5.5.1 Application for Dry Gas / Wet Gas ................................................................. 153 5.5.2 Application for Gas Condensate with Producible (Mobile) Reservoir Oil ..... 155

Chapter 6 Summary and Conclusions ...................................................................................... 158

Appendix A Input Data Summary .......................................................................................... 163 Appendix B User Guide .......................................................................................................... 170

vii

LIST OF FIGURES

Figure 2-1: Phase Diagram of Typical Gas Condensate Reservoir .......................................... 3

Figure 2-2: Distributions of Pseudo Components among Phases in Modified Black-Oil

Model ............................................................................................................................... 5

Figure 2-3: Graphical Representation of Zero-Dimensional Reservoir Model ........................ 7

Figure 2-4: Typical Field Performance of Gas Condensate – Gas and Oil Flow Rates vs.

Time ................................................................................................................................. 9

Figure 2-5: Typical Field Performance of Gas Condensate – Reservoir Pressure,

Bottomhole Flowing Pressure and Wellhead Pressure vs. Time ..................................... 10

Figure 2-6: Typical Field Performance of Gas Condensate – Cumulative Gas and Oil

Production vs. Time ......................................................................................................... 11

Figure 4-1: Graphical Representation of Standard PVT Properties ......................................... 44

Figure 4-2: Typical Characteristic of Gas Formation Volume Factor ( ) and Volatilized

Oil-Gas Ratio ( ) for Gas Condensate .......................................................................... 48

Figure 4-3: Typical Characteristic of Oil Formation Volume Factor ( ) and Solution

Gas-Oil Ratio ( ) for Gas Condensate ........................................................................... 48

Figure 4-4: Graphical Representation of CVD Data used in Walsh-Towler Algorithm .......... 53

Figure 4-5: Graphical Representation of Nodal Analysis ........................................................ 96

Figure 4-6: Graphical Representation of Field Optimization .................................................. 116

Figure 5-1: Simulated Gas Formation Volume Factor and Volatilized Oil-Gas Ratio of

Gas Condensate ................................................................................................................ 118

Figure 5-2: Simulated Oil Formation Volume Factor and Solution Gas-Oil Ratio of Gas

Condensate ....................................................................................................................... 119

Figure 5-3: Simulated Specific Gravity of Reservoir Gas ....................................................... 120

Figure 5-4: Volumes of Surface Gas Pseudo Component in Reservoir Gas Reservoir Oil,

and Cumulative Gas Production ....................................................................................... 121

Figure 5-5: Volumes of Stock-Tank Oil Pseudo Component in Reservoir Gas, Reservoir

Oil, and Cumulative Oil Production................................................................................. 122

Figure 5-6: Densities of Surface Gas and Stock-Tank Oil Pseudo Components at First

Stage Separator, Second Stage Separator and Stock Tank Condition .............................. 124

viii

Figure 5-7: Volumes of Stock-Tank Oil Pseudo Component in Reservoir Gas, Reservoir

Oil, and Cumulative Oil Production in term of Gas-Equivalent ...................................... 125

Figure 5-8: Total Volumes of Stock-Tank Oil Pseudo Component and Surface Gas

Pseudo Component in term of Gas-Equivalent ................................................................ 126

Figure 5-9: Simulated Production Results of Gas Condensate using Simplified Gas

Condensate Tank Model .................................................................................................. 130

Figure 5-10: Phase Envelope and Reservoir Depletion Paths at Two Different Reservoir

Temperatures .................................................................................................................... 133

Figure 5-11: Simulated Gas Formation Volume Factor and Volatilized Oil-Gas Ratio of

Volatile Oil using Gas Condensate PVT Model .............................................................. 134

Figure 5-12: Simulated Oil Formation Volume Factor and Solution Gas-Oil Ratio of

Volatile Oil using Gas Condensate PVT Model .............................................................. 134

Figure 5-13: Simulated Production Results of Volatile Oil using Simplified Gas

Condensate Tank Model .................................................................................................. 135

Figure 5-14: Mole Fraction Behavior of Vapor Phase Molar Fraction ( ) for Gas

Condensates and Volatile Oils ......................................................................................... 136

Figure 5-15: Cumulative Gas and Oil Production vs. Time..................................................... 138

Figure 5-16: Total Gas and Oil Flow Rates vs. Time .............................................................. 139

Figure 5-17: Reservoir Pressure, Bottomhole Flowing Pressure, and Wellhead Pressure

vs. Time ............................................................................................................................ 140

Figure 5-18: Gas Saturation and Specific Gravity of Reservoir Gas vs. Time ........................ 142

Figure 5-19: Total Gas Flow Rate ( ) vs. Cumulative Gas Production during Decline

Period .................................................................................................. 143

Figure 5-20: Decline Rate ( ) vs. Cumulative Gas Production during Decline Period

( ) ........................................................................................................... 145

Figure 5-21: Annual Expenditure, Annual Total Revenue, and Cumulative Discounted

Net Cash Flow vs. Production Time ................................................................................ 147

Figure 5-22: Net Present Value vs. Interest Rate ..................................................................... 148

Figure 5-23: Field Optimization Results .................................................................................. 149

ix

LIST OF TABLES

Table 4-1: Volume-Translation Coefficients for Pure Components (Whitson and Brule,

2000) ................................................................................................................................ 31

Table A-1: Pressures and Temperatures for Standard PVT Properties Calculation

Subroutine ........................................................................................................................ 163

Table A-2: Physical Properties of Pure Components ............................................................... 163

Table A-3: Binary Interaction Coefficients of Pure Components ............................................ 164

Table A-4: Volume Translation Coefficient of Pure Components .......................................... 164

Table A-5: Reservoir Input Data .............................................................................................. 165

Table A-6: Relative Permeability Input Data .......................................................................... 165

Table A-7: Standard PVT Properties ....................................................................................... 166

Table A-7: Standard PVT Properties (Cont.) ........................................................................... 167

Table A-8: Tubing Input Data .................................................................................................. 168

Table A-9: Economic Input Data ............................................................................................. 168

Table A-9: Economic Input Data (Cont.)................................................................................. 169

Table A-10: Field Performance Prediction Input ..................................................................... 169

Table A-11: Field Performance Optimization Input ................................................................ 169

x

NOMENCLATURE

Normal Symbol Definition

Reservoir drainage area

Hyperbolic decline exponent

Co-volume parameter of i-th component

Formation volume factor

Gas formation volume factor

Oil formation volume factor

Two-phase gas formation volume factor

Two-phase oil formation volume factor

Overall molar faction of i-th component

Formation (rock) compressibility

Deitz shape factor

Non-Darcy coefficient / Tubing Diameter

Decline rate

Expansivity of formation (rock)

Expansivity of reservoir gas

Expansivity of reservoir oil

Expansivity of reservoir water

Efficiency factor of tubing

Fanning’s friction factor

Fugacity of i-th component in vapor phase

Fugacity of i-th component in liquid phase

Moody’s friction factor

Molar fraction of vapor phase

Molar fraction of liquid phase at reservoir condition

Molar fraction of liquid phase

Molar fraction of liquid phase at first-stage separator

Molar fraction of liquid phase at first-stage separator produced from

reservoir gas

Molar fraction of liquid phase at first-stage separator produced from

reservoir oil

Molar fraction of liquid phase at second-stage separator

Molar fraction of liquid phase at second-stage separator produced from

reservoir gas

Molar fraction of liquid phase at second-stage separator produced from

reservoir oil

Molar fraction of liquid phase at stock-tank condition

Molar fraction of liquid phase at stock-tank condition produced from

reservoir gas

Molar fraction of liquid phase at stock-tank condition produced from

reservoir oil

Fugacity of i-th component in original fluid

xi

Fugacity of i-th component in liquid-like phase

Fugacity of i-th component in vapor-like phase

Amount of surface gas pseudo component / Gas in place

Amount of gas-equivalent pseudo component

Amount of surface gas pseudo component in reservoir gas phase

Amount of surface gas pseudo component in reservoir oil phase

Amount of cumulative gas injection

Amount of cumulative gas production

Cumulative gas production at year Cumulative gas production at abandonment condition

Cumulative gas production at end of plateau

Cumulative gas recovery

Incremental of cumulative gas production

Annual gas production at year Incremental of gas recovery

Reservoir thickness

Elevation of upstream node

Elevation of downstream node

Difference in elevation of downstream and upstream node

Absolute permeability of reservoir

Effective permeability of reservoir gas

Relative permeability of reservoir gas

Relative permeability of reservoir oil

Volatility ratio of i-th component

Tubing length

Temperature dependency coefficient of i-th component

Molecular weight of vapor phase

Molecular weight of gas at reservoir condition

Molecular weight of i-th component

Molecular weight of oil at reservoir condition

Molecular weight of oil at stock-tank condition

Molecular weight of oil at stock-tank condition produced from reservoir

gas

Molecular weight of oil at stock-tank condition produced from reservoir

oil

Molecular weight of liquid phase

Molecular weight of remaining fluid inside PVT cell

Number of component in multi-component hydrocarbon

Mole fraction of excess gas removed from PVT cell

Mole fraction of remaining gas inside PVT cell

Mole fraction of remaining gas inside PVT cell plus excess gas

Mole fraction of remaining oil inside PVT cell

Mole fraction of remaining fluid inside PVT cell

Amount of stock-tank oil pseudo component / Oil in place

Amount of stock-tank oil pseudo component in reservoir gas phase

Amount of stock-tank oil pseudo component in reservoir oil phase

xii

Amount of cumulative oil production

Cumulative oil production at year Cumulative oil production at abandonment condition

Cumulative oil recovery

Reynolds number

Incremental of cumulative oil production

Annual oil production at year Incremental of oil recovery

Original gas in place

Original oil in place

Pressure

Upstream pressure

Downstream pressure

Average pressure between upstream and downstream

Critical pressure of i-th component

Drawdown pressure inside the reservoir

Pseudocritical pressure

Reservoir pressure

Reservoir pressure at abandonment condition

Reservoir pressure at end of plateau

Reduced pressure of i-th component

Pressure at standard condition

Bottomhole flowing pressure

Bottomhole flowing pressure at end of plateau

Wellhead pressure

Minimum allowable wellhead pressure

Pressure drop from initial reservoir pressure

Total gas flow rate of the field

Total gas flow rate of the field at abandonment condition

Total gas flow rate of the field during plateau period

Total oil flow rate of the field

Total oil flow rate of the field at abandonment condition

Gas flow rate per well

Gas flow rate per well during plateau period

Annual average gas flow rate of the field

Annual average oil flow rate of the field

Reservoir radius

Wellbore radius

Universal gas constant

Gas-oil equivalent factor

Fugacity ratio of i-th component

Solution gas-oil ratio

Solution gas-oil ratio at bubble point pressure

Volatilized oil-gas ratio

Volatilized oil-gas ratio at dew point pressure

Target recovery factor at end of plateau

xiii

Fugacity ratio of i-th component in liquid-like phase Fugacity ratio of i-th component in vapor-like phase

Total skin factor

Volume-translate coefficient of i-th component

Mechanical skin factor

Average reservoir gas saturation

Minimum gas saturation

Sum of the mole number of liquid-like phase

Average reservoir oil saturation

Sum of the mole number of vapor-like phase

Average reservoir water saturation

Connate water saturation

Specific gravity of gas

Production time

Production time at abandonment condition

Production time at end of plateau

Temperature

Temperature of upstream node

Temperature of downstream node

Pipe section average temperature

Critical temperature of fluid

Critical temperature of i-th component

Pseudocritical temperature of

Reduced temperature of i-th component

Temperature at standard condition

Fluid velocity

Retrograde liquid volume fraction

Amount of excess gas at reservoir condition

Amount of remaining gas phase at reservoir condition

Amount of remaining gas phase plus excess gas at reservoir condition

Amount of remaining oil phase at reservoir condition

Pore volume of reservoir

Original volume of PVT cell

Molar volume of phase ―a‖ calculated from EOS

Critical molar volume of i-th component

Molar volume of vapor phase

Molar volume of vapor phase calculated from EOS

Molar volume of liquid phase

Molar volume of liquid phase calculated from EOS

Pseudocritical molar volume

Amount of water pseudo component in reservoir water

Amount of water influx

Amount of cumulative water injection

Amount of cumulative water production

Molar fraction of surface gas pseudo component in reservoir oil

Molar fraction of i-th component in liquid phase

Molar fraction of stock-tank oil pseudo component in reservoir oil

xiv

Molar fraction of surface gas pseudo component in reservoir gas

Molar fraction of i-th component in vapor phase

Molar fraction of stock-tank oil pseudo component in reservoir gas

Molar fraction of i-th component in liquid-like phase Molar fraction of i-th component in vapor-like phase

Molar number of i-th component in liquid-like phase Molar number of i-th component in vapor-like phase

Compressibility factor of fluid

Two-phase compressibility factor

Compressibility factor of phase ―a‖

Average compressibility factor

Greek Symbol Definition

Coefficient to adjust relative permeability of reservoir gas

Turbulence parameter

Specific gravity of gas

Binary interaction coefficient between i-th and j-th component

Tubing roughness

Fluid viscosity

Viscosity of vapor phase / Viscosity of reservoir gas

Viscosity of i-th component at low pressure

Viscosity of liquid phase

Viscosity of liquid phase at low pressure

Viscosity of reservoir oil

Fluid density

Density of vapor phase

Density of gas phase at reservoir condition

Density of liquid phase

Density of oil phase at reservoir condition

Density of oil phase at stock-tank condition

Density of oil phase at stock-tank condition produced from reservoir gas

Density of oil phase at stock-tank condition produced from reservoir oil

Pseudo reduced density of liquid phase

Density of remaining fluid inside PVT cell

Molar density of reservoir gas

Molar density of surface gas pseudo component

Molar density of reservoir oil

Molar density of stock-tank oil pseudo component

Annual production time

Average reservoir porosity

Fugacity coefficient of i-th component in vapor phase

Fugacity coefficient of i-th component

Fugacity coefficient of i-th component in liquid phase

Pitzer’s acentric factor of i-th component

xv

ACKNOWLEDGEMENTS

First and foremost I would like to thanks my advisor, Dr. Luis Ayala, for his continuous

guidance, support and friendship throughout my graduate study. Without his encouragement and

invaluable advice, this research would not have been completed. Additional thanks are extended

to Dr. Larry Grayson, and Dr. Li Li for their interest and time in serving as my thesis committee.

I would like to express my sincere appreciation to Dr. Turgay Ertekin and Dr. Russel

Johns, and Dr. Zuleima Karpyn for the fundamental knowledge they have taught. I am also very

grateful for educational environment that the faculty and staff of the Department of Energy and

Mineral Engineering have created. I highly thank my sponsor, PTT Exploration and Production

Company, for every support they have given.

Many friends and colleagues have been very supportive. I would like to express my

gratitude to Pipat Likanapaisal, Nithiwat Siripatrachai and Kanin Bodipat who always are good

friends throughout my student life at Pennsylvania State University. I also thank all of my

colleagues for making me have meaningful time and experience.

Finally, but most deeply, I am forever in dept to my family, my father Phiraphong

Vardcharragosad, my mother Pikun Tanarungreung, my sister Pungjai Keandoungchun, and

sister’s family, for their support, encouragement, and most importantly their tolerance.

Chapter 1

Introduction

Natural gas is a natural occurring gas which consisting of methane primarily. It plays a

significant role in global economic as one of the main sources of energy. In 2009, world natural

gas reserves equaled 6.29 Trillion Standard Cubic Feet (TCF) while world production reached

106 BCF for the year (EIA, 2011). Conventional reservoirs consist of five different fluid types:

dry gas, wet gas, retrograde gas, volatile oil, and black oils (McCain, 1990). They are

distinguished from each other based on the present of fluid phases inside the reservoir and at

surface production facilities.

Field development and investment decisions in petroleum and natural gas require an

integration of expertise from various areas including geology, reservoir, drilling, completion,

process, and economic. Location and size of reservoirs, production rates and time, total

recoverable volumes, number of wells and platforms, drilling and completion techniques,

processing facilities scheme, cost and revenue, etc. are examples of information required for

adequate field development decisions. Field performance indicators consist of information

regarding flow rates, pressures, and production time is very important for field development. If

field performance indicators are satisfactorily predicted, the hydrocarbon field could be

developed using the best possible exploitation strategy while optimizing its economic

performance. If not, the field might end up with too many wells, processing facilities that are too

large, or wrong equipment sizing which can jeopardize profits or even lead to significant losses of

investor’s capital.

In modern age, computer simulation is used to simulate various types of mathematical

models which can couple geological, fluid property, reservoir, production network, processing

2

facilities, and economic information. Field performance could be predicted by integrating these

models together. However, the required type of mathematical model needs to be carefully

selected to be able to perform the calculation most effectively. For reservoir characterization, for

example, the modeler might utilize either a fully dimensional numerical model - which can aptly

capture all reservoir heterogeneities and geometry by discretizing it into many small grids -, or a

zero-dimensional model - which assumes average reservoir and fluid properties across the

domain. For fluid behavior characterization, the modeler might select either a fully compositional

model based on the use of an equation of state and detailed fluid composition data, or a black-oil

model - which uses the pseudo-component concept and relies on PVT laboratory results.

Selection of those models generally depends on availability of input data, time constraint, and

required accuracy of simulation results. In this study, a zero dimensional model coupled with a

black-oil PVT fluid description is implemented for the study of field development optimization

strategies in retrograde natural gas reservoirs.

Chapter 2

Background

2.1 Gas Condensate Hydrocarbon Fluid

A gas condensate, retrograde gas condensate, or retrograde gas, is one of the five

reservoir fluid types (McCain, 1990). The typical phase envelope of gas condensate reservoirs is

shown in Figure 2-1. Gas condensates contain more intermediate and heavy hydrocarbon

components more than dry gases or wet gases. As shown in Figure 2-1, their reservoir

temperature is located in between the fluid’s critical temperature and their cricondentherm. The

reservoir depletion path of a gas condensate fluid typically crosses the dew point line and a liquid

phase appears at reservoir pressures lower than that of the dew point. The presence of liquid

phase in the reservoir significantly increases the system complexity, even if this liquid phase does

not flow and is very unlikely to be produced under normal production conditions.

Figure 2-1: Phase Diagram of Typical Gas Condensate Reservoir

Re

serv

oir

Pre

ssu

re

Reservoir Temperature

ReservoirDepletion

Path

SurfaceDepletion

Path

Critical Point

4

The general characteristics of gas condensate reservoir fluid can be summarized as

follows (Walsh and Lake, 2003):

Initial Fluid Molecular Weight: 23 – 40 lb/lbmol

Stock-Tank Oil Color: Clear to Orange

Stock Tank Oil Gravity: 45 – 60 API

C7-plus Mole Fraction: 0.01 – 0.12

Typical Reservoir Temperature: 150 – 300 F

Typical Reservoir Pressure: 1500 – 9000 psia

Volatilized Oil-Gas Ratio: 50 – 300 STB/MMSCF

Primary Recovery of Original Gas In Place: 70% – 85%

Primary Recovery of Original Oil In Place: 30% - 60%

5

2.2 Modified Black-Oil Model

A black-oil fluid model is a fluid characterization formulation which represents multi-

component hydrocarbon mixture in terms of two hydrocarbon pseudo components, namely the

―surface gas‖ and ―stock-tank oil‖ pseudo components. In a traditional black-oil model, the

solubility of the ―surface gas‖ pseudo component in the reservoir oil fluid phase is taken into

account while the solubility of ―stock-tank oil‖ pseudo component in reservoir gas phase is

neglected. The modified black-oil model which also called two-phase two-pseudo component

model does not neglect the ―stock tank oil‖ solubility in the gaseous reservoir phase, thus

including both solubility variables into the formulation. Figure 2-2 shows the distribution of

surface gas and stock-tank oil pseudo components among reservoir gas and reservoir oil phases.

Figure 2-2: Distributions of Pseudo Components among Phases

in Modified Black-Oil Model

The assumptions behind the modified black-oil PVT model can be summarized as

follows (Walsh and Lake, 2003 and Whitson and Brule, 2000):

There are two pseudo components which are surface gas and stock-tank oil.

There are two fluid phases which are reservoir gas (vapor) and reservoir oil

(liquid) phases.

Surface GasStock-Tank

Oil

Surface Gas Stock-Tank Oil

Reservoir GasPhase

Reservoir OilPhase

6

Surface gas pseudo component is reservoir fluid which remains in gas phase at

standard condition.

Stock-tank oil pseudo component is reservoir fluid which remains in oil phase at

standard condition.

The reservoir gas phase, which is reservoir fluid remains in vapor phase at

reservoir condition, consists of surface gas and stock-tank oil pseudo

components.

The reservoir oil phase, which is reservoir fluid remains in liquid phase at

reservoir condition, consists of surface gas and stock-tank oil pseudo

components.

Properties of surface gas and stock-tank oil pseudo components remain the same

throughout the reservoir depletion.

7

2.3 Zero-Dimensional Reservoir Model

The Material Balance Equation (MBE) is a specialized type of mass balance equation that

combines mass balance equations of all pseudo components present in the reservoir into single

equation. The MBE is also called zero-dimensional reservoir model or tank model because it

assumes that a reservoir behaves like a homogeneous tank with average rock and fluid properties

across the domain. Pressure, temperature, and compositional gradients are thus neglected. MBEs

can be derived from integrating diffusivity equations over space and time.

Figure 2-3: Graphical Representation of Zero-Dimensional Reservoir Model

(Source: http://www.joe.co.jp/english/menu2-5.html)

The following assumptions are implemented in traditional in zero-dimensional reservoir

models:

Reservoir is isothermal

Reservoir is under thermodynamic equilibrium condition

There are no chemical and biological reaction in reservoir

Capillary pressures of reservoir fluids are negligible

Gravitational gradients in reservoir are negligible

Pressure gradients in reservoir are negligible

8

2.4 Field Performance Prediction

A field performance prediction consists in the calculations of pressures, flow rates,

cumulative productions, and expected production times based on available reservoir, production

network, and production constraint data. Field life is divided into three periods which are build-

up, plateau, and decline periods (Ayala, 2009a), as depicted in Figure 2-4. During build-up

period, gas flow rate per well ( ) is kept constant while number of wells continuously

increases until total maximum number of wells needed for field development is reached. During

the plateau period, both gas flow rate per well ( ) and number of wells are fixed; therefore,

total gas flow rate ( ) (equal to gas flow rate per well ( ) times number of wells) remains

constant. During decline period, wellhead pressure ( ) is kept constant at the minimum

allowable wellhead pressure ( ). Under such conditions, reservoir pressure ( ) becomes too

low to maintain the target plateau rate, thus gas flow rate ( ) continuously declines until

abandonment condition is reached.

9

Figure 2-4: Typical Field Performance of Gas Condensate

– Gas and Oil Flow Rates vs. Time

Figure 2-4 through Figure 2-6 show the typical field performance predictions for the

development of a gas condensate reservoir. As indicated earlier, gas flow rate ( ) increases

with increasing production time during the build-up period because more wells are put on

production. Then, it is kept constant until the end of the plateau period. During final decline

period, gas flow rate ( ) continuously decreases with production time because reservoir

pressure ( ) becomes not enough to sustain the plateau rate. Above dew point conditions, oil

flow rate ( ) produced at the surface becomes directly proportional to gas flow rate ( ).

However, once reservoir conditions reach the dew point, condensate production at the surface

becomes a function of both gas flow rate ( ) and the volatilized oil-gas ratio ( ) at reservoir

conditions. Figure 2-4 shows that even if gas flow rate ( ) is maintained at a constant target

value during the plateau period, oil flow rate ( ) can actually decreases because of decreasing

volatilized oil-gas ratio ( ) below the dew point.

qo

sc-

Tota

l Oil

Flo

w R

ate

qg

sc-

Tota

l Gas

Flo

w R

ate

Production Time

qgsc

qosc

Dew Point

Build-up Plateau Decline

10

Figure 2-5: Typical Field Performance of Gas Condensate

– Reservoir Pressure, Bottomhole Flowing Pressure

and Wellhead Pressure vs. Time

Figure 2-5 demonstrates that, during field development calculations, reservoir pressure

( ) decreases as production time increases because more oil and gas are being removed from the

reservoir. Wellhead pressure ( ) is also continuously decreased in time in order to maintain the

gas flow rate ( ) per well during the build-up and plateau periods. After that, once wellhead

pressure ( ) reaches the minimum allowable wellhead pressure at surface conditions, the

plateau gas flow rate ( ) cannot be maintained any longer and the decline period starts.

Bottomhole flowing pressure ( ) changes along with changes in reservoir pressure ( ) and

wellhead pressure ( ) in order to provide the required pressure drop within the reservoir and

production tubing.

Pre

ssu

re

Production Time

pr

pwf

pwh

Build-up Plateau Decline

Minimum Allowable Wellhead Pressure

11

Figure 2-6: Typical Field Performance of Gas Condensate

– Cumulative Gas and Oil Production vs. Time

Figure 2-6 shows cumulative gas ( ) and cumulative oil ( ) production are directly

related to their corresponding flow rates. Above dew point conditions, both of them increase at

the same pace. Below the dew point, however, cumulative oil production ( ) builds up at a

much slower rate compared to that of cumulative gas production ( ) because of the increased

reservoir condensation driven by a decreasing volatilized oil-gas ratio ( ). Recovery factor of

gas at abandonment condition would therefore become much higher than the recovery factor of

oil or condensate because large amounts of condensate are left behind as immobile phase inside

the reservoir.

Np

-C

um

ula

tive

Oil

Pro

du

ctio

n

Gp

-C

um

ula

tive

Ga

s P

rod

uct

ion

Production Time

Gp

Np

Build-up Plateau Decline

12

2.5 Visual Basic for Applications (VBA)

Visual Basic for Applications (VBA) is a programming language from Microsoft. The

program is built into most MS-Office applications i.e. MS-Word, MS-Excel, MS-Access. Users

can use VBA to create calculation subroutine and control user interface features such as menus,

toolbars, worksheets, charts, etc (Walkenbach, 2007). VBA can only run within the host

application, and not as a standalone application. VBA is functionally rich, and flexible. Because it

is built into MS-Office applications, VBA subroutines will be able to execute so long as those

applications are available on computer machines. MS-Excel with built in VBA is a very favorable

platform for developing simulations. The main reasons are that most of engineers are familiar

with MS-Excel application and MS-Excel itself is user-friendly software with many useful built-

in features. Excel’s worksheets could be used as table to store input data. Simulation results could

be easily stored in the tabular form and displayed on various types of built-in chart.

Chapter 3

Problem Statement

In the development of a petroleum and natural gas reservoir, projected field performance

is the most important information required by all relevant people involved in the process of

design, risk assessment, and decision-making process. Field performance analysis can require a

significant amount of expertise and time, especially for more complex reservoir fluid system such

as gas condensates. The use of the appropriate modeling approach is the key to analyze the field

performance most efficiently. Full scale, fully dimensional, commercial simulators might not be

able to yield the best or optimized solutions even if they are based on of highly sophisticated

mathematical models. This is because more sophisticated and detailed models are subject to the

availability of very detailed set of reservoir and fluid data, which is typically scarce, and time

constraints and demands. In the analysis of gas condensates, for example, commercial simulators

often rely on compositional modeling for fluid property calculations. Compositional models can

accurately simulate reservoir fluid properties; however, it is sophisticated model and can take a

relatively long time to run. For reservoir fluid flow characterization, commercial simulators

generally rely on fully dimensional numerical models which could perfectly capture reservoir

heterogeneities; yet, they can take a significant amount of time to construct, conceptualize, and

execute. Again, the limited availability and uncertainty of required input data such as fluid

composition, reservoir heterogeneities, capillary pressures, and relative permeabilities could

significantly impact the reliability of the results obtained from these sophisticated models.

This study aims at developing a model which can efficiently and inexpensively perform

field performance analysis and optimization tasks for gas condensate reservoirs. The proposed

model utilizes a zero-dimensional reservoir formulation coupled with a pseudo component or

14

black oil PVT formulation for fluid properties calculation. These models are relatively simple, but

fast, reliable, and robust. Results show that the proposed model is able to predict field

performance while faithfully capturing the most salient characteristics of gas condensate

reservoirs. In addition, optimization on targeted variables can be accomplished without difficulty.

Chapter 4

Model Description

The proposed field performance predictor has been developed using Microsoft Excel with built-in

Visual Basic for Applications (VBA) subroutines. Workflow begins with the simulation of

standard black-oil PVT properties, which could be done either based on standard PVT laboratory

results (such as the Constant Volume Expansion or CVD) or via a phase behavior model based on

cubic equations of state. Next, field performance data is calculated by integrating a zero-

dimensional reservoir model, standard PVT fluid properties, well performance models for flow

rates and pressure calculation, and production constraints. Based on this, an economic analysis

can be performed based on simplified economic model. Finally, optimization on target variables

can be carried out by evaluating field performance and net present value repeatedly for different

and plausible production scenarios. The proposed simulation tool has been designed to simulate a

single gas condensate reservoir based on the continuous drilling of identical wells placed at

different locations of the reservoir area. Wellhead pressure is used to control gas flow rate.

Reservoir pressure is used as the abandonment criteria. Optimization variables are target recovery

factor at end of plateau and total number of wells. Those variables could be re-selected by simple

modification in the VBA code. However, optimization variables can be made independent for a

real field operation.

16

4.1 Phase Behavior Model (PBM)

A phase behavior calculation (or a flash calculation) is used to predict the phase behavior

of a reservoir fluid at an equilibrium condition. A standard phase behavior model consists of four

main calculation modules; namely, compressibility factor calculations, vapor-liquid equilibrium

calculations, fluid properties predictions, and phase stability analysis, which must be fully

integrated to perform the flash calculation. The calculation starts with the determination of

number of co-existent phases or phase stability analysis. If fluid is found in a single phase (stable)

condition, fluid properties are calculated based on the available information on overall fluid

composition. If fluid is found in a two-phase (unstable) condition, composition and molar fraction

of each phase are determined using vapor-liquid equilibrium calculations. Then properties of each

co-existent phase are calculated based on fluid composition of that phase (Ayala, 2009b).

Input data consists of pressure, temperature, overall composition, physical properties,

binary interaction coefficients, and volume translation coefficient of each pure component. Peng-

Robinson Equation-of-State (PR EOS) is used to calculate Pressure-Volume-Temperature (PVT)

relationship of the reservoir fluid (Peng and Robinson, 1976). Vapor-liquid equilibrium is

assumed and an overall species material balance for a two-phase system is enforced. The output

from a PBM subroutine consists of number of phases, molar fraction, composition, molecular

weight, compressibility factor, density, adjusted density and viscosity of each fluid phase.

17

4.1.1 Compressibility Factor

Compressibility factor or Z-factor is volumetric multiplier utilized to convert ideal gas

volumes, as predicted by the ideal gas equation of state, to real gas volumes, as realized

experimentally. Compressibility factor is a fundamental and very important variable because

other fluid properties can be calculated based on compressibility factor data. Z-factor calculation

subroutine is developed based on generalized formulation (Coats, 1985). Although Peng-

Robinson EOS is utilized throughout this study, other EOSs could also be applied by

implementing simple modifications outlined below

When the fluid is in a single phase condition, overall composition will be inputted into

generalized formula for the calculation of the single-phase compressibility factor. However, when

the fluid is in two-phase condition, composition of each phase must be first calculated based on

vapor-liquid equilibrium calculations in order to estimate the corresponding compressibility

factors of each phase.

Generalized Formulation

Compressibility factor depends on the chosen PVT relationship or equation of state

(EOS). The generalized formula for cubic EOS proposed by Coats is utilized (Coats, 1985). This

form can be applied for Redlich-Kwong (RK), Soave-Redlich-Kwong (SRK), and Peng-Robinson

(PR) EOSs (Redlich, O. and Kwong, J.N.S. 1949, Soave, G. 1972, and Peng and Robinson,

1976).

18

Equation 4-1

where:

= number of components in the multi-component hydrocarbon

= molar fraction of the i-th component

= binary interaction coefficient between the i-th and j-th components

= reduced pressure of the i-th component =

= reduced temperature of the i-th component =

= critical pressure of the i-th component {psia}

= critical temperature of the i-th component {R}

= pressure {psia}

= temperature {R}

19

, which accounts for the temperature dependency built into the molecular attraction

parameter, is calculated from Equation 4-2 for PR EOS and from Equation 4-3 for SRK EOS.

Equation 4-2

Equation 4-3

where:

= Pitzer’s acentric factor of the i-th component

Pressure, temperature, molar fraction, and properties of pure components are input into

the generalized EOS formula shown above which yields a cubic polynomial in Z. Analytical,

semi-analytical, or numerical approach can be used to solve this cubic equation. In this work, the

analytical approach is applied.

20

Z-Factor Selection

Because of the nature of cubic equation, more than one root could be found for any given

pressure, temperature, and fluid composition. As described by Danesh (p. 176), the following

criteria are used for Z-factor selection (Danesh, 1998). If there is only one real root, Z-factor is

equal to that root. If there is more than one real root, the following criteria must be applied.

The intermediate root will always be rejected.

If the minimum Z-factor is less than B, maximum Z-factor will be selected.

If the minimum Z-factor is higher than B, The root that provides the lower Gibbs

energy will be selected.

Z-factor which is less than B must be rejected because when Z-factor is less than B,

molar volume becomes smaller than the co-volume. For this reason, such Z-factor would have no

physical meaning. For the last condition in the list above, Equation 4-4 is used to find the root

with lower Gibbs energy. Following Danesh (1998), if the right hand side of this equation is

positive, minimum Z-factor will be selected. Otherwise, the maximum Z-factor will be selected.

Equation 4-4

21

4.1.2 Vapor-Liquid Equilibrium

Two main components are considered in order to predict properties of multi-component

hydrocarbon in Vapor-Liquid Equilibrium (VLE) condition: material balance considerations and

thermodynamic considerations. Iterative procedure is applied until the solution that satisfies both

criteria can be determined.

Material Balance Considerations

Rachford and Rice objective function, which is derived from enforcing an overall species

mass balance in a two-phase multi-component system, is utilized to calculate molar fraction of

each phase (Rachford and Rice, 1952):

Equation 4-5

where:

= molar faction of i-th component

= volatility ratio of i-th component =

= molar fraction of i-th component in vapor phase

= molar fraction of i-th component in liquid phase

= molar fraction of vapor phase

22

After solving for from the objective function, molar fraction of liquid phase is

calculated from Equation 4-6, composition of vapor phase is calculated from Equation 4-7, and

composition of liquid phase is calculated from Equation 4-8.

Equation 4-6

Equation 4-7

Equation 4-8

Thermodynamic Considerations

According to the second law of thermodynamics, any system in equilibrium, such as a

VLE condition, must have the maximum possible entropic state under the prevailing conditions.

For such condition to be established, thermodynamics shows that net transfer of heat, momentum,

and mass between both phases must be zero. Thus, temperature, pressure, and every species

chemical potential in both phases must be equal to each other.

23

Chemical potential cannot be measured directly. However, equality of chemical potential

can be represented by equality of fugacity between both phases. Fugacity is the pressure

multiplier to correct non-ideality and to make ideal gas equation work for real gas during Gibbs

energy calculations. In a VLE condition, fugacity of liquid phase must be equal to fugacity of

vapor phase. Equation 4-9 is used to calculate fugacity for vapor phase while Equation 4-10 is

used for liquid phase.

Equation 4-9

Equation 4-10

where:

= fugacity of i-th component in vapor phase

= fugacity of i-th component in liquid phase

= fugacity coefficient of i-th component in vapor phase

= fugacity coefficient of i-th component in liquid phase

= molar fraction of i-th component in vapor phase

= molar fraction of i-th component in liquid phase

= pressure {psia}

24

For the generalized formula of cubic EOSs discussed above, fugacity coefficients can be

calculated using Equation 4-11 (Coats, 1985) below. Definitions of parameters are the same as

definitions used in Equation 4-1. It should be noted that is equal to for calculating fugacity

coefficient of a liquid phase and is equal to for calculating fugacity coefficient of a vapor

phase.

Equation 4-11

Volatility ratio ( ) is equal to ratio between the gas composition and the liquid

composition during an equilibrium condition. For a system with a VLE condition, is equal

to . By substituting Equation 4-9 and Equation 4-10 into definition of volatility ratio, volatility

ratio can be expressed in terms of fugacity coefficients as follows.

Equation 4-12

25

The Successive Substitution Method

From material balance consideration, molar fraction of vapor phase and composition of

each phase are functions of volatility ratios and overall composition. Volatility ratios themselves

are also function of composition of each phase. Thus, an iterative procedure is needed in order to

perform VLE prediction and honor the fugacity equality constraint. The following procedure is

used to perform two-phase flash calculation (Whitson and Brule, 2000, p.52-55).

First, initial guesses of volatility ratios are calculated using Equation 4-13 as proposed by

Wilson (Wilson, 1968). Rachford and Rice objective function (Equation 4-5) is then solved using

a standard Newton-Raphson iterative method. Then, the compositions of each phase are

calculated using Equation 4-7 and Equation 4-8.

Equation 4-13

Next, the fugacity values of each component in both liquid and vapor phases are

calculated using Equation 4-9 through Equation 4-11. Successive Substitution Method (SSM) is

utilized to update volatility ratios (Equation 4-14) for a next iteration as shown below

Equation 4-14

26

where:

= volatility ratio of i-th component at iteration level n

= fugacity of i-th component in liquid phase at iteration level n

= fugacity of i-th component in vapor phase at iteration level n

Once volatility ratios are updated, convergence criteria presented in Equation 4-15 must

be checked. If the criteria are not satisfied, the procedure is repeated by solving Rachford and

Rice objective function and recalculating phase compositions and resulting fugacities until

convergence is attained.

Equation 4-15

The SSM algorithm is expected to have slow convergence rate near the critical point. To

avoid this problem, accelerated SSM algorithm has been proposed. The algorithm proposed by

Michelsen (Michelsen, 1982b) or the algorithm proposed by Merah et al (Merah et al, 1983) are

examples of well-known ASSM algorithms.

27

4.1.3 Fluid Property Prediction

Molecular Weight

Molecular weight of vapor and liquid phases are weighted average of molecular weight

of all pure components, as shown below

Equation 4-16

Equation 4-17

where:

= molecular weight of vapor phase {lb/lbmol}

= molecular weight of liquid phase {lb/lbmol}

= molecular weight of i-th component {lb/lbmol}

= mole fraction of i-th component in vapor phase

= mole fraction of i-th component in liquid phase

= number of components in the multi-component hydrocarbon

28

Density

Density of each phase is calculated from Equation 4-18 and Equation 4-19.

Equation 4-18

Equation 4-19

where:

= density of vapor phase {lbm/ft3}

= density of liquid phase {lbm/ft3}

= molecular weight of vapor phase {lbm/lbmol}

= molecular weight of liquid phase {lbm/lbmol}

= molar volume of vapor phase {ft3/lbmol}

= molar volume of liquid phase {ft3/lbmol }

29

Molar volume of each phase is calculated from real gas law (Equation 4-20), then,

adjusted by using volume-translation technique.

Equation 4-20

where

= calculated molar volume of phase ―a‖ from EOS {ft

3/lbmol}

= compressibility factor of phase ―a‖

= universal gas constant {10.732 psi-ft3/R-lbmol}

= temperature {R}

= pressure {psia}

As discussed by Whitson and Brule (p.51) and Danesh (p.141-143), calculated molar

volume from real gas law can be adjusted by implementing volume-translation or volume-shift

technique (Whitson and Brule, 2000 and Danesh, 1998). This technique improves volumetric

calculation of liquid phase, which is the main problem of two-constant EOS’s, without altering

VLE prediction results. The volume translation technique, originally introduced by Martin and

further developed by Penelous et al and Jhaveri and Youngren, can be summarized as follows

(Martin, 1979, Penelus et al, 1982, and Jhaveri and Youngren, 1988):

30

Calculated molar volumes from the selected EOS are corrected by using Equation 4-21

and Equation 4-22. Component-dependent volume-shift parameters ( ) are calculated from

Equation 4-23 and volume-translate coefficients are in Table 4-1.

Equation 4-21

Equation 4-22

where:

= corrected molar volume of liquid phase

= corrected molar volume of vapor phase

= calculated molar volume of liquid phase from EOS

= calculated molar volume of vapor phase from EOS

= component-dependent volume-shift parameter

= molar fraction of i-th component in liquid phase

= molar fraction of i-th component in vapor phase

= number of components in the multi-component hydrocarbon

31

Equation 4-23

where:

= component-dependent volume-shift parameter

= co-volume parameter of i-th component

= volume-translate coefficient of i-th component

Table 4-1: Volume-Translation Coefficients for Pure Components (Whitson and Brule, 2000)

Component PR EOS SRK EOS

N2 -0.1927 -0.0079

CO2 -0.0817 0.0833

H2S -0.1288 0.0466

C1 -0.1595 0.0234

C2 -0.1134 0.0605

C3 -0.0863 0.0825

i-C4 -0.0844 0.0830

n-C4 -0.0675 0.0975

i-C5 -0.0608 0.1022

n-C5 -0.0390 0.1209

n-C6 -0.0080 0.1467

n-C7 0.0033 0.1554

n-C8 0.0314 0.1794

n-C9 0.0408 0.1868

n-C10 0.0655 0.2080

32

Viscosity

Viscosity of vapor phase is calculated from the correlation proposed by Lee et al in 1966

(Equation 4-24 through Equation 4-27).

Equation 4-24

Equation 4-25

Equation 4-26

Equation 4-27

where:

= viscosity of vapor phase {cp}

= density of vapor phase {lbm/ft3}

= molecular weight of vapor phase {lbm/lbmol}

= temperature {R}

33

The viscosity of a liquid phase is calculated from the correlation proposed by Lohrenz et

al in 1964. The correlation is originally proposed by Jossi et al in 1962 for calculating viscosity

of pure component. Lohrenz et al extend the use of original correlation to hydrocarbon mixtures.

It should be noted that the formula in Lohrenz et al’s paper contains a typing error on coefficient

0.040758 for the cubic density term.

Equation 4-28

where:

= viscosity of liquid phase {cp}

= viscosity of liquid phase at low pressure {cp}

= viscosity parameter of liquid phase (mixture) {cp-1

}

= pseudo reduced density of liquid phase

Viscosity of liquid phase at low pressure is calculated from Equation 4-29, Equation

4-30, and Equation 4-31. A conversion factor of 5.4402 is used to convert original units (K and

atm) to oil field units (R and psia).

Equation 4-29

34

Equation 4-30

Equation 4-31

where:

= viscosity of liquid phase at low pressure {cp}

= molar fraction of i-th component in liquid phase

= viscosity of i-th component at low pressure {cp}

= viscosity parameter of i-th component {cp-1

}

= reduce temperature of i-th component ( )

= temperature {R}

= critical temperature of i-th component {R}

= critical pressure of i-th component {psia}

= molecular weight of i-th component {lbm/lbmol}

= number of components

Viscosity parameter of liquid phase is calculated from Equation 4-32 to Equation 4-35.

Equation 4-32

35

Equation 4-33

Equation 4-34

Equation 4-35

where:

= viscosity parameter of liquid phase (mixture) {cp-1

}

= pseudocritical temperature of liquid phase {R}

= critical temperature of i-th component {R}

= pseudocritical pressure of liquid phase {psia}

= critical pressure of i-th component {psia}

= molecular weight of liquid phase {lbm/lbmol}

= molecular weight of i-th component {lbm/lbmol}

= molar fraction of i-th component in liquid phase

= number of components

36

Pseudo reduced density of the liquid phase is calculated from Equation 4-36 and

Equation 4-37 shown below.

Equation 4-36

Equation 4-37

where:

= pseudo reduced density of liquid phase

= density of liquid phase {lbm/ft3}

= molecular weight of liquid phase {lbm/lbmol}

= pseudocritical molar volume of liquid phase {ft3/lbmol}

= critical molar volume of i-th component {ft3/lbmol}

= molar fraction of i-th component in liquid phase

= number of components

37

4.1.4 Phase Stability Analysis

The ability to predict whether the system is in single phase (stable) or multiple phases

(unstable) is crucial in a VLE or flash calculation. Whitson and Brule (p.55-61) discuss the

graphical representation as well as numerical algorithm of phase stability analysis based on the

studies by Baker et al and Michelsen (Whitson and Brule, 2000; Baker et al, 1982; Michelsen,

1982a). These studies explain how the Gibbs tangent-plane criteria can effectively be used to

analyze the phase stability problem. The phase stability analysis subroutine utilized by this study

has been developed based on these calculation procedures, which can be summarized in the 11

steps outlined below.

Step 1: Calculate the mixture fugacity from overall composition using Equation 4-9 /

Equation 4-10 and Equation 4-11. The Z-factor yielding the lowest Gibbs energy should be

utilized for the calculation of mixture fugacity.

Step 2: Use Wilson’s equation to estimate initial values (Equation 4-13).

Step 3: Calculate second-phase mole number, , using the mixture composition and the

estimated K values.

Equation 4-38

Equation 4-39

38

where:

= mole number of i-th component in vapor-like phase

= mole number of i-th component in liquid-like phase

= mole fraction of i-th component

= volatility ratio of i-th component

Step 4: Sum the mole numbers of vapor-like phase ( ) and liquid-like phase ( ).

Equation 4-40

Equation 4-41

Step 5: Normalize the mole numbers to get the mole fraction of i-th component in vapor-

like phase, and liquid-like phases,

Equation 4-42

Equation 4-43

39

Step 6: Calculate the fugacity of vapor-like and liquid-like phases based on the calculated

mole fraction from Step 5. Equation 4-9 , Equation 4-10 and Equation 4-11 are utilized.

Step 7: Calculate the fugacity ratio corrections for successive substitution update of the

values.

Equation 4-44

Equation 4-45

where:

= fugacity ratio calculation of i-th component in vapor-like phase

= fugacity ratio calculation of i-th component in liquid-like phase

= fugacity of i-th component in original fluid

= fugacity of i-th component in vapor-like phase

= fugacity of i-th component in liquid-like phase

= Sum the mole numbers of vapor-like phase

= Sum the mole numbers of liquid-like phase

40

Step 8: Check whether convergence criteria is achieved

Equation 4-46

Step 9: If convergence is not obtained, update values

Equation 4-47

Step10: Apply criterion to check whether a trivial solution has been obtained

Equation 4-48

Step 11: If a trivial solution is not indicated, go to Step 3 for the next iteration.

41

The following criteria are used to interpret the results from this numerical algorithm:

If the tests on both vapor-like and liquid-like phases satisfy trivial solution

criterion, the system of interest is stable (single phase)

If sum of the mole numbers on both vapor-like and liquid-like phases is less than

or equal to 1.0, the system of interest is stable (single phase).

If one of the pseudo phases satisfies trivial solution criterion and sum of the mole

numbers of the other pseudo phase is less than or equal to 1.0, the system is

stable (single phase).

Otherwise, the system is unstable; both vapor and liquid phases coexist.

42

4.2 Standard PVT Properties

The standard PVT properties used to describe a two-phase, two-pseudo component fluid

model (―black oil model‖) relies on the definition and calculation of four basic properties,

namely: gas formation volume factor ( ), oil formation volume factor ( ), volatilized oil-gas

ratio ( ), and solution gas-oil ratio ( ). These PVT properties are required inputs for a zero-

dimensional reservoir model. In this study, these required PVT properties can be obtained from

either a laboratory fluid analysis, typically a Constant Volume Depletion (CVD) test, or from a

phase behavior model (PBM) calculation. If the PVT/CVD laboratory report is available, the

resulting PVT properties are calculated using Walsh-Towler algorithm (Walsh and Lake, 2003).

A template has been prepared using MS-Excel worksheet for this purpose. In the absence of a

PVT lab report, a PBM calculation is implemented which combines Walsh-Tolwer method with

the work of Thararoop in 2007 (Thararoop, 2007). This PBM subroutine does not only extend the

flexibility of the main simulator significantly, but also provide very useful information about fluid

properties which could help in thoroughly analyzing the depletion characteristics of the given gas

condensate fluid.

The specific gravity of reservoir gas is required for flow rate and flowing pressure

calculations, as it will be discussed below. The specific gravity of a reservoir gas can be obtained

from either the laboratory fluid analysis or from molecular weight calculations derived from

PBM. If the lab analysis is available, compositions of the produced wellstreams reported in the

experimental depletion study based on the Constant Volume Depletion (CVD) test are used to

calculate molecular weight of reservoir gas. If the lab report is unavailable, the molecular weight

of the reservoir gas is obtained directly from flash/PBM calculation results. Specific gravity of

reservoir gas is equal to molecular weight of reservoir gas divided by molecular weight of air.

43

4.2.1 Definitions, Mathematic Relationships, and Characteristics

A clear understanding of the definitions of standard PVT ―black oil‖ properties that are

used to characterize two-phase, two-pseudo component fluid models is crucial for their

meaningful calculation and prediction. These definitions, mathematic relationships, and their

most significant features have been summarized below (Walsh and Lake, 2003; Whitson and

Brule, 2000).

Definitions

Figure 4-1shows the graphical representation of the definitions of the standard PVT

properties used in the formulation of two-phase, two-pseudo component fluid model (or modified

―black-oil‖ model). In this figure, the gas phase at reservoir condition ( ) results from the mixing

of certain amounts of surface gas ( ) and stock-tank oil ( ) pseudo components. The oil

phase at reservoir condition ( ) results from the mixing of certain amounts of surface gas ( )

and stock-tank oil ( ) pseudo components. The produced gas phase at surface condition ( )

(not shown in the figure) would consists of the combination of surface gas pseudo component

produced from gas phase at reservoir condition ( ) and surface gas pseudo component liberated

from oil phase at reservoir condition ( ). By the same token, the produced oil phase at surface

condition ( ) (not shown in the figure) consists of stock-tank oil pseudo component produced

from oil phase at reservoir condition ( ) and stock-tank oil pseudo component condensed from

gas phase at reservoir condition ( ).

44

Figure 4-1: Graphical Representation of Standard PVT Properties

Based on the pseudo component definitions described above, the definitions of the

associated ―black oil‖ properties can be straightforwardly presented. For example, the formation

volume factor for the gas ( ) would be basically defined as ratio between volume of gas phase at

reservoir condition ( ) and volume of surface gas pseudo component produced from that

reservoir gas, evaluated at surface conditions ( ). Formation volume factor of oil ( ) is

defined as ratio between volume of oil phase at reservoir condition ( ) and volume of stock-tank

oil pseudo component produced from that reservoir oil, evaluated at surface condition ( ).

Volatilized oil-gas ratio ( ) is defined as ratio between volume of stock-tank oil ( ) and

volume of surface gas ( ) pseudo components produced from the same reservoir gas ( ),

evaluated at surface condition. Solution gas-oil ratio ( ) is defined as ratio between volume of

surface gas ( ) and volume of stock-tank oil ( ) pseudo components produced from the same

reservoir oil ( ), evaluated at surface condition. Mathematically, Equation 4-49 through

Vg

Vo

PR, TR

Reservoir Condition Surface Condition

Gfg

Nfg

Gfo

Nfo

Reservoir Gas

Reservoir Oil

Surface Gas

Stock-Tank Oil

Vg

Gfg

Bg = ----------- Rv = -----------

Nfg

Gfg

Bo = ----------- Rs = -----------

Vo

Nfo

Gfo

Nfo

Psc, Tsc

45

Equation 4-52 summarize, in oil field units, the standard PVT properties based on these

definitions and the nomenclature presented in Figure 4-1.

Equation 4-49

Equation 4-50

Equation 4-51

Equation 4-52

It follows from the preceding discussion that reservoir fluid compositions can be

calculated for the envisioned pseudo binary mixture. For example, the molar fraction of surface

gas pseudo component in the gas phase at reservoir conditions, defined as , should be directly

related to the value of Rv. Molar fraction of stock-tank oil pseudo component in gas phase at

reservoir condition would be defined as . Clearly, + = 1. For the oil reservoir phase, the

molar fraction of surface gas pseudo component in the oil phase at reservoir conditions would be

, and should be directly related to the value of Rs The molar fraction of stock-tank oil pseudo

46

component in oil phase at reservoir condition is thus defined as . Clearly, + = 1. Their

formulas are summarized in Equation 4-53 through Equation 4-56.

Equation 4-53

Equation 4-54

Equation 4-55

Equation 4-56

Mathematic Relationships

If only one mole of reservoir fluid is considered, volumes at reservoir condition, and

, can be represented by molar density at reservoir condition, and , respectively.

Similarly, volumes at surface condition, , , , and , can be represented by molar

fraction of pseudo component in reservoir fluid and molar density at surface condition, ,

, , and , respectively. If we substitute these definitions into equations for

47

standard PVT properties and substitute densities of gases with real gas equation, the following

expressions can be derived.

Equation 4-57

Equation 4-58

Equation 4-59

Equation 4-60

Depletion Characteristics

Figure 4-2 and Figure 4-3 show the typical depletion behavior of the standard PVT

properties for the case of a gas condensate reservoir fluid. Similar behavior can be found in the

work by Walsh and Lake (Walsh and Lake, 2003, p.493) for the case of field-data derived

properties.

48

Figure 4-2: Typical Characteristic of Gas Formation Volume Factor ( )

and Volatilized Oil-Gas Ratio ( ) for Gas Condensate

Figure 4-3: Typical Characteristic of Oil Formation Volume Factor ( )

and Solution Gas-Oil Ratio ( ) for Gas Condensate

Rv

-V

ola

tiliz

ed

Oil

-Gas

Rat

io

Bg

-G

as F

orm

atio

n V

olu

me

Fac

tor

Reservoir Pressure

Dew Point Pressure

Rv

Bg

Rs

-So

luti

on

Gas

-Oil

Rat

io

Bo

-O

il Fo

rmat

ion

Vo

lum

e F

acto

r

Reservoir Pressure

Dew Point Pressure

Bo

Rs

49

As shown in Figure 4-2, gas formation volume factors ( ) are expected to increase with

decreasing reservoir pressure ( ) because the denominator, , in Equation 4-57 approaches zero.

Volatilized oil-gas ratio ( ) will remain constant because all parameters in Equation 4-59 remain

the same. Constant values of , , and result from the constant composition of gas phase in

the reservoir. Once dew point conditions are reached, Figure 4-2 also shows that the volatilized

oil-gas ratio ( ) is expected to decrease with decreasing reservoir pressure, mainly because of

decreasing and increasing values in Equation 4-59. Driven by the condensate drop out that

develops in the reservoir below dew point conditions, the reservoir gas will start to contain less

heavy hydrocarbon molecules that can be produced as condensate at surface condition. As a

result, the fraction of stock-tank oil ( ) in the reservoir gas decreases while fraction of surface

gas ( ) increases ( + = 1). As pressure depletion progresses, and if it gets low enough, the

volatilized oil-gas ratio ( ) trend would be reversed.

Figure 4-3 illustrates that at reservoir pressure above the dew point there is no liquid

phase at reservoir condition and therefore no calculations of and can be directly performed

from their definitions. Once dew point conditions are crossed, oil formation volume factor ( ) is

expected to decrease with decreasing reservoir pressure mainly because of increasing and

values in Equation 4-58. As pressure decreases, more surface gas pseudo component will be

liberated from the oil phase. As a result, the molar fraction of stock-tank oil pseudo component in

oil phase ( ) becomes higher and the density of oil phase at reservoir condition ( ) also

increases. Similarly, the solution gas-oil ration ( ) will be expected to decrease with decreasing

reservoir pressure because of the increased molar fraction of stock-tank oil pseudo component in

oil phase ( ) and decreasing molar fraction of surface gas pseudo component in oil phase ( ) in

Equation 4-60. Even though oil formation volume factors ( ) and solution gas-oil ratios ( )

cannot be calculated directly because of the lack of an actual liquid phase at reservoir from their

50

definitions, Walsh and Lake suggest employing the following relationships for oil formation

volume factor ( ) and solution gas-oil ratio ( ) as ―place-holder‖ values above the dew point:

Equation 4-61

Equation 4-62

51

4.2.2 Obtaining Standard PVT Properties from Laboratory PVT Reports

In a laboratory PVT test, a representative sample of the reservoir fluid is subjected to a

series of depletion steps that try to closely mimic or reproduce the expected pressure depletion

path followed by the fluid during reservoir production. Temperature of the test is maintained

constant and equal to prevailing reservoir temperature. Resulting volumes of each phase (liquid

and vapor) are recorded along with the pressure at which the record is made. Fluid composition

and physical properties of the produced fluids are also analyzed. The typical standardized PVT

tests conducted for gas condensate fluids are the Constant Composition Expansion (CCE) and

Constant Volume Depletion (CVD) tests. Details of these PVT tests can be found in many

petroleum engineering textbooks (McCain, 1990; Denesh, 1998; Whitson and Brule, 2000, Walsh

and Lake, 2003); thus, they will be discussed very briefly in this manuscript.

In a CCE test, the reservoir fluid sample is placed inside a PVT cell and is pressurized to

a pressure equal to initial reservoir pressure, while maintaining a constant temperature inside the

PVT cell equal to reservoir temperature. Pressure inside the cell is then decreased to a next lower

pressure level by isothermal expansion. The new volume of each phase is recorded. This process

continues until abandonment pressure conditions are reached. In the CCE testing process, no fluid

is taken out the cell and therefore the overall composition of reservoir fluid inside the PVT cell

remains constant while the volumes and densities of each the co-existing phases below dew point

conditions do change with cell pressure.

In a CVD test, a reservoir fluid sample will be placed inside the PVT cell and pressurized

to the dew point pressure, while the temperature of the PVT cell is kept constant at reservoir

temperature. Then, pressure of the cell will be lowered to the next pressure level by isothermal

expansion. After that, a portion of gas phase inside the cell is produced (i.e., removed out of the

cell) so that the cell’s volume is restored back to the original cell volume at dew point conditions.

52

The volume that the liquid phase occupies inside the PVT cell is recorded and the excess

(produced) gas analyzed. Depletion study which provides the resulting cumulative production

data at every pressure level is recorded and is used during the calculation of the standard PVT

properties from laboratory PVT fluid test report.

In this study, a calculation template is prepared in MS-Excel worksheet. The Walsh-

Towler algorithm is implemented to convert the results from the CVD experiments into the

standard table of PVT properties for a gas condensate fluid. Walsh-Towler algorithm is

summarized below.

Walsh-Towler Algorithm

Walsh-Towler algorithm is one of the methods used to calculate standard PVT properties

for gas condensate based on CVD testing results (Walsh and Towler, 1995; Walsh and Lake,

2003). This algorithm is relatively simple because it based on enforcing material balance

constraints around the PVT cell at every pressure level during the PVT lab test. The algorithm

was originally proposed by Walsh and Towler in 1995 and was later modified by Walsh and Lake

in 2003. By directly using data from a CVD report, this algorithm is implicitly assuming that

actual field separator conditions of the surface production system is the same as those surface

condition used during the CVD PVT test. It also assumes that only the gas phase at reservoir

condition can be recovered and that any condensate drops out inside the reservoir will remain

immobile during reservoir life.

One of the constraints of using this method is the availability of cumulative production

data at surface conditions because such data is not always performed or reported for every CVD

experiment. If such cumulative production data at surface conditions is not available in the CVD

report, it is customarily recommended to implement surface flash calculations using Standing’s

53

K-values to reproduce them (Walsh and Lake, 2003). The algorithm also requires a high accuracy

and reliability of the CVD report in order to obtain a healthy and physically meaningful set of

derived standard PVT properties. It can be demonstrated that small error in the data reported by a

CVD test can result in PVT property values which are physically impossible (e.g., negative

values). And even when the data reported by the CVD report is highly reliable, the Walsh and

Towler algorithm can still lead to unphysical values for standard PVT properties. This limitation

results from combining the two-phase two-pseudo component (―black oil‖) model with material

balance calculation around the PVT cell. This limitation will be discussed in detail in Chapter 5.

Walsh-Towler algorithm consists of six sequential steps which must be fully completed at

every given pressure level before moving to the next pressure. One pre-calculation is also needed

before starting the algorithm. The variables and their nomenclature employed in the sequence of

calculations are graphically illustrated in Figure 4-4.

Figure 4-4: Graphical Representation of CVD Data used in Walsh-Towler Algorithm

Vg,j

Vo,j

Reservoir Condition Surface Condition

Gfg,j

Nfg,j

Gfo,j

Nfo,j

Reservoir Gas

Reservoir Oil

Surface Gas

Stock-Tank Oil

Vg,j

Gfg,j

Bg = ----------- Rv = -----------

Nfg,j

Gfg,j

Bo = ----------- Rs = -----------

Vo,j

Nfo,j

Gfo,j

Nfo,j

VT

VEG,j

PR ≥ PDew PR < PDew ∆Gpj

∆Npj

54

Pre-calculation: In this step, the total cumulative volumes of surface gas ( ) and stock-

tank oil ( ) pseudo components produced from the reservoir fluid, and the resulting volume of

PVT cell ( ) are calculated for the dew point condition. The volume of surface gas pseudo

component ( ) is calculated from the summation of cumulative gas recovery from 1st stage

separator, 2nd

stage separator, and stock tank for all available pressures - from dew point

conditions to the last reported (abandonment) pressure. The volume of stock-tank oil pseudo

components ( ) is equal to cumulative oil recovery from stock tank for all available and reported

pressures (dew point to abandonment). These data are obtained from the calculated cumulative

recovery reported in the depletion table.

PVT cell’s volume is calculated from the definition of gas formation volume factor

(Equation 4-63). The gas formation volume factor ( ) is calculated from Equation 4-57.

Compressibility factor of gas phase ( ) can be obtained from the CVD report. Mole fraction of

surface gas pseudo component in the reservoir gas ( ) is equal to divided by the volume of gas

equivalent at the dew point ( ) which is usually taken as 1000 MSCF.

Equation 4-63

Volatilized oil-gas ratio at dew point ( ) is calculated from Equation 4-64, while oil

formation volume factor ( ) and solution gas-oil ratio ( ) are calculated from Equation 4-61

and Equation 4-62, respectively.

55

Equation 4-64

Step 1: Find and : Starting at the dew point, the volume of surface gas pseudo

component released from the excess gas ( ) at each pressure is calculated from the summation

of cumulative gas recovery from 1st stage separator, 2

nd stage separator, and stock tank. Volume

of and stock-tank oil pseudo component released from the same excess gas ( ) at each pressure

is equal to cumulative oil recovery from stock tank. These data are obtained from the calculated

cumulative recovery reported in the depletion table. Incremental of and from pressure level

j-1 to pressure level j are calculated from Equation 4-65 and Equation 4-66. Please note that

pressure level j begins from zero at the dew point (j=0). , , , and are also

equal to zero.

Equation 4-65

Equation 4-66

Step 2: Find and : Total volume of surface gas ( ) and stock-tank oil ( ) pseudo

components released from both reservoir gas and reservoir oil at pressure level j are calculated

from Equation 4-67 and Equation 4-68. It should be noted that pressure level j begins from zero at

the dew point (j=0), and and are equal to and , respectively.

56

Equation 4-67

Equation 4-68

Step 3: Find and : Volume of oil phase at reservoir condition at pressure level j

( ) is calculated from Equation 4-69. Retrograde liquid volume fraction at pressure level j

( ), can be obtained from CVD report. Volume of gas phase after excess gas removal at

reservoir condition at pressure level j ( ) is calculated from Equation 4-70. Note that pressure

level j begins at zero at dew point conditions (j=0)

Equation 4-69

Equation 4-70

Step 4: Find , , and : Molar fraction of reservoir fluid which remains in the

PVT cell at pressure level j ( ) is calculated from Equation 4-71. For this calculation, two-

phase compressibility factor ( ) data can be obtained from the CVD report. Molar fraction of

excess gas which is removed from PVT cell at pressure level j ( ) is calculated from Equation

4-72. Molar fraction of gas phase which remain in PVT cell at pressure level j ( ) is calculated

from Equation 4-73. Compressibility factor of gas ( ) is also obtained from the CVD report.

57

Please note that pressure level j begins from zero (j=0) at the dew point. and at dew

point are equal to 1.0 while at dew point is equal to zero.

Equation 4-71

Equation 4-72

Equation 4-73

Step 5: Find and : Volume of surface gas pseudo component produced from

reservoir gas at pressure level j ( ) is calculated from Equation 4-74. Volume of stock-tank

pseudo component produced from reservoir gas at pressure level j ( ) is calculated from

Equation 4-75. It is important to note that pressure level j begins from zero at the dew point (j=0).

and at dew point pressure are equal to and , respectively.

Equation 4-74

58

Equation 4-75

Step 6: Find and : Volume of surface gas pseudo component produced from

reservoir oil at pressure level j ( ) is calculated from Equation 4-76. Volume of stock-tank oil

pseudo component produced from reservoir oil at pressure level j ( ) is calculated from

Equation 4-77.

Equation 4-76

Equation 4-77

After completing all six steps outline above for the given pressure level, Equation 4-49

through Equation 4-52 are now directly used to calculate the standard PVT properties. All

applicable unit conversion factors must be checked and adjusted properly. The calculation

process is systematically repeated for all pressure levels until all reported data in the CVD report

have been considered and abandonment conditions have been reached.

Standard PVT properties at pressures higher than the dew point are calculated based on

the properties at dew point pressure. Gas formation volume factor ( ) is the product of gas

formation volume factor at dew point pressure and relative volume obtained directly from CCE

testing results. The relative volume is the ratio between total volume of hydrocarbon at reservoir

conditions and the volume at saturated conditions. For under-saturated gas condensate system,

59

relative volume is equal to the ratio between at specified pressure and at dew point pressure.

Volatized oil-gas ratio ( ) is equal to volatilized oil gas ratio at dew point pressure. Oil

formation volume factor ( ) and solution gas-oil ratio ( ) are calculated from Equation 4-61

and Equation 4-62, respectively.

Finally, it is very important to mention that, in Walsh-Towler algorithm, volumes of

pseudo components produced from the reservoir oil (step 6) do not actually come from direct

surface measurement. In a CVD test, the oil inside the cell is never produced (is assumed

immobile) so surface data for produced oil is not available.. Instead, these values are indirectly

calculated based on the enforcement of mass balance constraints around the PVT cell. Therefore,

actual oil formation volume factor ( ) and solution gas-oil ratio ( ) calculated from actual

surface flashes of the reservoir fluid might be significantly different from the ones estimated

using these indirectly calculated surface volumes. If the calculated and resulting from the

application of this algorithm do not agree with the physically acceptable trends or values, the

results should be disregarded and the laboratory results have to be adjusted.

60

4.2.3 Obtaining Standard PVT Properties from a Phase Behavior Model

Another method for simulating standard PVT properties for gas condensate is to utilize

Phase Behavior Model (PBM). This method is based on combination of the algorithm used in

Walsh-Towler method and the work of Thararoop in 2007. The general idea of this method is to

substitute CVD testing results with the outputs from flash calculation. Mass balance around PVT

cell, which is used to obtain the properties of reservoir oil in Walsh-Towler algorithm, is replaced

with an actual flash calculation performed for both the reservoir gas and oil phases. Chapter 5

will discuss about the impact from these changes in more detail.

Input data required for this method include initial reservoir condition, surface separator

conditions, initial reservoir fluid composition, physical properties, binary interaction coefficients,

and volume translation coefficients of pure components. The simulation algorithm consists of

nine calculation steps and a pre-calculation. Parameters used in those equations were represented

graphically in Figure 4-4.

Pre-calculation: First, dew point pressure is determined using a phase stability

calculation. Then, mole of initial reservoir fluid inside PVT cell ( ), volume of PVT cell ( ),

volume of surface gas ( ) and stock-tank oil ( ) pseudo components are evaluated at dew point

condition.

The dew point pressure is determined by performing Phase Stability Analysis. Stability of

initial reservoir fluid is continuously evaluated at different pressure levels, while temperature is

controlled at reservoir temperature. Pressure level starts at initial reservoir pressure; then, it is

continuously decreased by 1.0 psi interval until the initial reservoir fluid becomes unstable. The

last pressure level that initial reservoir fluid is in stable condition is the dew point pressure. A

61

direct calculation of saturation pressure at the prevailing reservoir temperature could be also

alternatively employed (Whitson and Brule, 2000).

The initial amount of mole of the reservoir fluid sample inside PVT cell ( ) is

calculated from Equation 4-78. Standard condition is set to be 14.7 psia and 520 R. Volume of

initial reservoir fluid in term of gas equivalent ( ) is assumed to be 1.0 MMSCF which is used

as the basis for the calculation.

Equation 4-78

The associated volume of PVT cell ( ) is calculated from Equation 4-79. Molecular

weight ( ) and density ( ) are obtained by performing flash calculation on initial reservoir

fluid composition at the dew point condition.

Equation 4-79

The molar fractions of surface gas ( ) and stock-tank oil ( ) pseudo components in

reservoir fluid are calculated from Equation 4-80 and Equation 4-81. Molar fraction of liquid

phase at first-stage separator (

) is obtained by performing flash calculation on initial

reservoir fluid composition at first-stage separator condition. Molar fraction of liquid phase at

second-stage separator (

) is obtained by performing flash calculation on liquid composition

62

from first-stage separator at second-stage separator condition. Molar fraction of liquid phase at

stock-tank condition ( ) is obtained by performing flash calculation on liquid composition

from second-stage separator at stock-tank condition.

Equation 4-80

Equation 4-81

Total volume of surface gas ( ) and stock-tank oil ( ) pseudo components initially

present in the reservoir fluid are calculated from Equation 4-82 and Equation 4-83. Value of

379.56 is molar volume of gases at standard condition which is constant. Molecular weight

( ) and density (

) of oil at stock-tank condition are obtained from flash calculation

results at stock-tank condition. Please note that these values (G and N) are not being obtained by

cumulative adding cumulative production values at every pressure level, as done in the original

Walsh and Tower algorithm. Chapter 5 will present a discussion on this regard and justification.

Equation 4-82

Equation 4-83

63

The gas formation volume factor ( ) is calculated from Equation 4-84. Volatilized oil-

gas ratio at dew point ( ) is calculated from Equation 4-64 by implementing the proper unit

conversion factor. Oil formation volume factor ( ) and solution gas-oil ratio ( ) are calculated

from Equation 4-61 and Equation 4-62, respectively.

Equation 4-84

Step 1: Find and : Moles of gas phase present at reservoir conditions before

the removal of excess gas at every pressure level j ( ) is calculated from Equation 4-85.

Moles of oil phase remaining at reservoir conditions at pressure level j ( ) is calculated from

Equation 4-86. Molar fraction of gas phase at reservoir condition at pressure level j ( ) is

obtained from performing flash calculation on overall composition from pressure level j-1, at

pressure level j. Note that pressure level j begins from zero (j=0) at the dew point. is

equal to , is equal to zero, and is equal to .

Equation 4-85

Equation 4-86

64

Step 2: Find and : The volume that the gas phase occupies at reservoir

condition before the removal of the excess gas at every pressure level j ( ) is calculated

from Equation 4-87. The volume of reservoir oil phase present at pressure level j ( ) is

calculated from Equation 4-88. Molecular weight and density of gas and oil phases at reservoir

condition at pressure level j ( ,

, ,

) are obtained by performing flash

calculation on overall composition from pressure level j-1, at pressure level j. Note that pressure

level j begins from zero at the dew point (j=0). is equal to and is equal to zero.

Equation 4-87

Equation 4-88

Step 3: Find and : The volume of reservoir gas phase after excess gas removal at

pressure level j ( ) is calculated from Equation 4-89. Volume of excess gas at reservoir

condition at pressure level j ( ) is then calculated from Equation 4-90.

Equation 4-89

65

Equation 4-90

Step 4: Find and : Remaining moles of gas phase at reservoir condition after

excess gas removal at every pressure level j ( ) is calculated from Equation 4-91. Moles of

excess gas which are removed at pressure level j ( ) is then calculated from Equation 4-92.

Density and molecular weight are the same as those in Equation 4-87.

Equation 4-91

Equation 4-92

Step 5: Find and : The molar fractions or compositions of surface gas ( ) and

stock-tank oil ( ) pseudo components in the reservoir gas at every pressure level j are

calculated from Equation 4-93 and Equation 4-94. The fraction of liquid phase at first-stage

separator recovered from reservoir gas at pressure level j (

) is obtained by performing flash

calculation on composition of reservoir gas at pressure level j, at first-stage separator condition.

The fraction of liquid phase at second-stage separator recovered from reservoir gas at pressure

level j (

) is obtained by performing flash calculation on liquid composition from first-stage

66

separator at second-stage separator condition. The fraction of liquid phase at stock-tank condition

recovered from reservoir gas at pressure level j ( ) is obtained by performing flash

calculation on liquid composition from second-stage separator at stock-tank condition.

Equation 4-93

Equation 4-94

Step 6: Find and : Volume of surface gas ( ) and stock-tank oil ( )

pseudo components in reservoir gas at pressure level j are calculated from Equation 4-95 and

Equation 4-96. The value of 379.56 is molar volume of gases at standard condition which is a

constant for ideal gases. Molecular weight ( ) and density (

) of oil at stock-tank

condition recovered from reservoir gas at pressure level j are obtained from flash calculation

results at stock-tank condition in Step 5.

Equation 4-95

Equation 4-96

67

Step 7: Find and : The molar fractions of surface gas ( ) and stock-tank oil

( ) pseudo components in the reservoir oil at every pressure level j are calculated from

Equation 4-97 and Equation 4-98. The fraction of liquid phase at first-stage separator recovered

from reservoir oil at pressure level j (

) is obtained by performing flash calculation on

composition of reservoir oil at pressure level j, at first-stage separator condition. The fraction of

liquid phase at second-stage separator recovered from reservoir oil at pressure level j (

) is

obtained by performing flash calculation on liquid composition from first-stage separator at

second-stage separator condition. The fraction of liquid phase at stock-tank condition recovered

from reservoir oil at pressure level j ( ) is obtained by performing flash calculation on liquid

composition from second-stage separator at stock-tank condition.

Equation 4-97

Equation 4-98

Step 8: Find and : The volume of surface gas ( ) and stock-tank oil ( )

pseudo components in reservoir oil at pressure level j are calculated from Equation 4-99 and

Equation 4-100. The value of 379.56 is molar volume of gas at standard condition which is

constant. Molecular weight ( ) and density (

) of oil at stock-tank condition recovered

from reservoir oil at pressure level j are obtained from flash calculation results at stock-tank

condition in Step 7.

68

Equation 4-99

Equation 4-100

Step 9: Find and : Remaining moles of reservoir fluid inside PVT cell at pressure

level j ( ) is calculated from Equation 4-101. Overall composition of i-th component insider

PVT cell at pressure level j ( ) after gas removal is updated by implementing Equation 4-102.

Note that pressure level j begins from zero (j=0) at the dew point. is equal to . Liquid

composition ( ) and vapor composition ( ) of i-th component at pressure level j are obtained

by performing flash calculation on overall composition from pressure level j-1, at pressure level j.

Equation 4-101

Equation 4-102

After completing all nine steps outlined above at every given pressure level, Equation

4-49 through Equation 4-52 will be used to directly calculate standard PVT properties. All unit

69

conversion factors must be checked and properly adjusted. This calculation process must be

continuously repeated for the every pressure level until abandonment pressure is reached.

Standard PVT properties at pressures higher than the dew point are calculated based on

available properties at dew point pressure. Gas formation volume factor ( ) is calculated from

gas formation volume factor at dew point pressure using Equation 4-103. The ratio between

( ) at dew point pressure and ( ) at specified pressures above the dew point is equivalent to

ratio between volume of reservoir gas ( ) at specified pressures above the dew point and volume

of reservoir gas ( ) at dew point pressure. Volatized oil-gas ratio ( ) is equal to volatilized oil-

gas ratio at dew point pressure ( ). Oil formation volume factor ( ) and solution gas-oil ratio

( ) are calculated from Equation 4-61 and Equation 4-62, respectively.

Equation 4-103

70

4.3 Zero-Dimensional Reservoir Model

The Material Balance Equation (MBE) (also known as zero-dimensional reservoir model

or tank model) is a mass balance statement that combines mass balance equations of all pseudo

components present in the reservoir fluid. The assumptions behind a tank model have been

already addressed in Section 2.3. Walsh and Lake (2003) have presented a generalized form of

material balance equation that could be used for the analysis of depletion performance for all five

types from reservoir fluids, based on the work originally published by Walsh (1995). They also

developed the MBE specialized for gas condensate fluids by simplifying the generalized MBE for

the conditions particular to these kind of fluids. Section 4.3.1 discusses and presents the GMBE

proposed by Walsh as implemented in this study.

In zero-dimensional reservoir model, cumulative productions of pseudo components and

saturations of reservoir fluids are calculated as functions of reservoir pressure, standard PVT

properties, and initial reservoir condition. This model treats a reservoir as a homogeneous tank;

thus only average reservoir pressure and average PVT properties are required as the model inputs.

In this study, a VBA subroutine has been developed to simulate cumulative oil and gas

productions as well as their saturations as a function of reservoir pressure, by implementing the

MBE specialized for gas condensate fluids. Most of the time, the MBE is used to simulate the

results explicitly as a function of time and depletion. However, if some target outputs are

specified, such as cumulative recovery at end of plateau, an iterative procedure would need to be

implemented in order to honor the additional constraint.

71

4.3.1 Generalized Material Balance Equation

Generalized Material Balance Equation (GMBE) is the most generalized form of Material

Balance Equation which can be applied to all types of reservoir fluids. Walsh and Lake derived

the GMBE by combining mass balance equation of pseudo components, surface gas, stock-tank

oil and stock-tank water, with the saturation constraint and standard PVT properties described in

section 4.2 (Walsh and Lake, 2003). The following assumptions are assumed in addition to the

general assumptions for zero-dimensional reservoir model.

Reservoir consists of surface gas, stock-tank oil, and stock-tank water pseudo

components

Reservoir consists of gas, oil, and water phases.

Surface gas pseudo component is in reservoir gas and oil phases.

Stock-tank oil pseudo component is in reservoir gas and oil phases.

Stock-tank water pseudo component is in reservoir water phase.

Surfaces gas, stock-tank oil, and stock-tank water can be produced

Surface gas and stock-tank water can be injected into the reservoir

Water phase can enter into reservoir by water influx from aquifer

GMBE can be manipulated into many different forms. One of the most useful forms of

the GMBE is shown in Equation 4-104. The terms on the left-hand side represents net reservoir

expansion terms while terms on the right-had side represents net reservoir withdrawal. Net

reservoir expansion consists of net reservoir gas expansion, net reservoir oil expansion, net

reservoir water expansion, net formation expansion, and water influx. Net reservoir withdrawal

consists of net gas and oil withdrawal, and net water withdrawal.

72

Equation 4-104

where:

= volume of surface gas pseudo component in reservoir gas at initial

condition {SCF}

= volume of stock-tank oil pseudo component in reservoir oil at

initial condition {STB}

= volume of water component in reservoir water at initial condition

{STB}

= pore volume at initial condition {RB}

= volume of water influx {RB}

= cumulative gas production {SCF}

= cumulative gas injection {SCF}

= cumulative oil production {STB}

= cumulative water production {STB}

= cumulative water injection {SCF}

= expansivity of reservoir gas {RB/SCF}

= expansivity of reservoir oil {RB/STB}

= expansivity of reservoir water {RB/STB}

= expansivity of formation (rock) {Dimensionless}

= gas formation volume factor {RB/SCF}

= oil formation volume factor {RB/STB}

73

= water formation volume factor {RB/STB}

= volatilized oil-gas ratio {STB/SCF}

= solution gas-oil ratio {SCF/STB}

Expansivity of reservoir fluid is defined as the total expansion of a unit mass of reservoir

fluid between two reservoir pressures at the same reservoir temperature. Expansivities of

reservoir gas, reservoir oil, and reservoir water are calculated from Equation 4-105, Equation

4-106, and Equation 4-107, respectively. Expansivity of formation (rock) is defined in a different

form from fluid expansivity and is calculated in terms of formation (rock) compressibility as

indicated by Equation 4-108.

Equation 4-105

Equation 4-106

Equation 4-107

Equation 4-108

74

where:

= two-phase gas formation volume factor {RB/SCF}

= two-phase oil formation volume factor {RB/SCF}

= formation (rock) compressibility {psi-1

}

= pressure drop from initial reservoir pressure {psi}

The two-phase formation volume factor implemented above is defined as the ratio

between total volume of reservoir fluid (gas and oil phases) and total volume of the pseudo

component. Two-phase formation volume factor of gas ( ) and oil ( ) are calculated from

Equation 4-109 and Equation 4-110, respectively. If reservoir is a single phase gas reservoir, two-

phase gas formation volume factor ( ) will be equal to gas formation volume factor ( ) while

two-phase oil formation will remain undefined. Similarly, if reservoir is single phase oil reservoir,

two-phase oil formation volume factor ( ) will be equal to oil formation volume factor ( )

while the two-phase gas formation volume factor will remain undefined.

Equation 4-109

Equation 4-110

75

4.3.2 Material Balance Equation for a Gas Condensate Fluid

GMBE can be simplified significantly when condensate drop out, developed below dew

point saturation conditions in the reservoir, is considered immobile. The immobile condensate

assumption is a fairly reasonable one for gas condensates; however, it cannot be applied for other

types of reservoir fluid (Walsh and Lake, 2003). The Simplified Gas Condensate Tank model,

SGCT, is derived from Generalized Material Balance Equation with the following additional

assumptions:

Reservoir is under-saturated at initial reservoir pressure

Expansivities of water and formation are negligible

There is no water influx, water production, and water injection

There is no gas injection

Condensate drop out in the reservoir is immobile

Gas Condensate Performance Below Dew Point

At initial undersaturated conditions, the volume of surface gas pseudo component in

reservoir gas at initial condition is equal to the Original Gas In Place (OGIP or G) while and the

volume of stock-tank oil pseudo component in reservoir oil at initial condition is equal to zero.

Equation 4-111 is the SGCT model after applying all these additional assumptions:

Equation 4-111

76

This SGCT model can be further manipulated in order to obtain a more useful form by

dividing it through by and substituting by . After that, finite difference

approximation is applied, resulting in expressions for the calculation of incremental oil and gas

production. As a result, Equation 4-112 through Equation 4-118 are a set of equations that can be

used to calculate reservoir performance from SGCT model.

Equation 4-112

Equation 4-113

Equation 4-114

Equation 4-115

Equation 4-116

77

Equation 4-117

Equation 4-118

where

= incremental gas recovery from pressure level to

= incremental oil recovery from pressure level to

= cumulative gas recovery at pressure level

= cumulative oil recovery at pressure level

= gas formation volume factor at pressure level {RB/SCF}

= oil formation volume factor at pressure level {RB/STB}

= volatilized oil-gas ratio at pressure level {STB/SCF}

= solution gas-oil ratio at pressure level {SCF/STB}

= two-phase gas formation volume factor at pressure level

{RB/SCF}

= two-phase oil formation volume factor at pressure level

{RB/SCF}

78

It is important to note that this set of equations (Equation 4-112 through Equation 4-118)

only applies for pressure below the dew point. Pressure level j begins at the first pressure below

the dew point. and at pressure level j-1 are cumulative results from the calculation

above the dew point. The standard PVT properties are either obtained from any of the procedures

described in section 4.2 or are given as input data. Two-phase formation volume factors are

calculated from Equation 4-109 and Equation 4-110. Calculation of SGCT model must be fully

completed at one pressure level before moving onto the next pressure level.

Gas Condensate Performance Above Dew Point

At pressure higher than dew point pressure, SGCT model should be further modified by

substituting and into Equation 4-111. As a result, the MBE for an

under-saturated gas condensate collapses to the typical MBE for a wet gas, shown in Equation

4-119. A set of equations which mimic the calculation procedure of SGCT model below dew

point can be developed by using a finite difference approach. Resulting performance prediction

equations are shown in Equation 4-120 to Equation 4-123.

Equation 4-119

Equation 4-120

79

Equation 4-121

Equation 4-122

Equation 4-123

It should be noted that this set of equation (Equation 4-120 to Equation 4-123) only

applies for pressures above dew point. The pressure level j begins from zero at initial reservoir

and end at the last pressure above the dew point. , , , and are equal to

zero at the initial reservoir pressure. The calculation has to be completed at one pressure level

before moving to the next pressure level.

80

4.3.3 Phase Saturation Calculations

One of the methods to derive a phase saturation equation is to combine the mass balance

equation for the stock-tank oil pseudo-component with the saturation equation constraint in order

to eliminate gas saturation parameter ( ) and then combine the resulting

equation and volumetric OGIP calculation equation so that the pore volume variable is

eliminated. The resulting saturation equation for the oil phase is shown in Equation 4-124.

Equation 4-124

where:

= average reservoir oil saturation

= average reservoir gas saturation

= average reservoir water saturation

= cumulative oil production {STB}

= original oil in place {STB}

= average reservoir porosity

= gas formation volume factor {RB/SCF}

= oil formation volume factor {RB/STB}

= volatilized oil-gas ratio {STB/SCF}

= subscript for initial condition

81

For the typical gas condensate reservoir, initial oil saturation ( ) is equal to zero

because its initial pressure is typically found above dew point conditions. If formation/rock

expansion is neglected, porosity of the porous medium would remain constant at the value of

initial porosity ( ). If net water withdrawal, influx, and expansion are neglected, water

saturation would remain constant at the value of connate water saturation ( ).

Thus, for such conditions, Equation 4-124 can be significantly simplified to Equation 4-125.

Equation 4-125

For saturation calculation of gas condensate reservoir, if reservoir pressure is equal to or

higher than dew point pressure, average reservoir oil saturation is equal to . Otherwise,

average reservoir oil saturation is calculated from Equation 4-125.

82

4.3.4 Volumetric OGIP/OOIP Calculations

Original Gas In Place ( ) and Original Oil In Place ( ) can be calculated using

volumetric method. Equation 4-128 and Equation 4-129 are used to calculate OGIP and OOIP,

respectively, based on the premise that both surface gas and stock-tank oil pseudo components

can be found in the reservoir gas and reservoir oil phases. In these equations, multiplication of

drainage area, reservoir thickness, and porosity represent the pore volume of the reservoir. The

conversion factor of 7758 is used to convert ―acre-ft‖ unit to ―RB‖ unit. The first and the second

terms inside the bracket of Equation 4-126 represent volumes of surface gas pseudo component in

reservoir oil and reservoir gas phases, per reservoir pore volume, respectively. The first and the

second terms inside the bracket of Equation 4-127 represents volume of stock-tank oil pseudo

component in reservoir oil and reservoir gas phases, per reservoir pore volume, respectively.

Equation 4-126

Equation 4-127

where:

= reservoir drainage area {acre}

= reservoir thickness {ft}

= reservoir porosity

83

= initial oil saturation

= initial gas saturation

= oil formation volume factor at initial condition {RB/STB}

= gas formation volume factor at initial condition {RB/SCF}

= solution gas-oil ratio at initial condition {SCF/STB}

= volatilized oil-gas ratio at initial condition {STB/SCF}

For gas condensate reservoir which do not initially have an oil phase ( ), Equation

4-126 and Equation 4-127 can be simplified into Equation 4-128 and Equation 4-129,

respectively. Reservoir properties which are drainage area ( ), thickness ( ), porosity ( ), and

initial water saturation ( ) must be known. Fluid properties which are initial gas formation

volume factor ( ) and initial volatilized oil-gas ratio ( ) are obtained from standard PVT

property estimations.

Equation 4-128

Equation 4-129

84

4.4 Flow Rates and Flowing Pressures Calculation

In this study, flow rates and flowing pressures calculations are based on the

implementation of Inflow Performance Relationships (IPR) and Tubing Performance

Relationships (TPR). The IPR equation relates flow rates from reservoir into the wellbore with

the difference between reservoir pressure and bottomhole flowing pressure. The TPR equation

relates wellbore flow rates between a bottomhole to a surface location in terms of the difference

between bottomhole flowing pressure and wellhead pressure.

Because of the immobile condensate assumption, oil flow rates can be calculated as a

function of gas flow rates and volatilized oil-gas ratio. Moreover, because a pipeline equation

based on the homogeneous flow assumption (single pseudo phase) can be used as the TPR

equation, by implementing an appropriate tubing efficiency factor, gas flow rate is the only

parameter that needs to be determined. In field performance prediction, either the desired gas

flow rate or the target wellhead pressure will be specified. If gas flow rate is specified, the IPR

equation will be used to explicitly calculate bottomhole flowing pressure, and the TPR equation

will be used to explicitly calculate wellhead pressure. If wellhead pressure is specified,

bottomhole flowing pressure and gas flow rate are simultaneously solved by using nodal analysis

method to determine the solution of the IPR and TPR system of equations. In this study, two

subroutines have been developed for gas flow rate and bottomhole flowing pressure calculations

based on IPR equations. Similarly, two subroutines have been developed for gas flow rate and

wellhead pressure calculations based on TRP equations. For the numerical solution, nodal

analysis method is also implemented by using bisection iterative procedure.

85

4.4.1 Inflow Performance Relationship (IPR)

The pseudo steady state (PSS) flow rate from the reservoir into the wellbore of radius

flow in cylindrical-shape reservoir, closed boundary can be calculated using Equation 4-130 (Lee

et al, 2003).

Equation 4-130

For other reservoir shapes, PSS flow rates can be calculated by applying shape factor

concept (Deitz, 1965). Flow rate calculation with Deitz shape factor ( ) is shown in Equation

4-131. The Deitz shape factor ( ) is equal to 31.62 for a circular drainage area with a well

located at the center of the reservoir, and equal to 30.88 for a square drainage area with a well

located at the center of the reservoir.

Equation 4-131

For the two-phase two-pseudo component model, the surface gas pseudo-component is

recovered from both reservoir gas and reservoir oil (as dissolved gas); while stock-tank oil pseudo

component is recovered from reservoir oil and reservoir gas (as volatilized oil). By introducing

the concept of phase mobilities and phase relative permeabilities, Equation 4-131 can then be

modified to calculate flow rates for this two-phase two-pseudo component model. Gas flow rate

86

( ) and oil flow rate ( ) can be calculated from Equation 4-132 and Equation 4-133,

respectively.

Equation 4-132

Equation 4-133

where:

= flow rate of surface gas {SCF/D}

= flow rate of stock-tank oil {STB/D}

= absolute permeability of reservoir {md}

= relative permeability of reservoir gas

= relative permeability of reservoir oil

= average reservoir pressure {psia}

= bottomhole flowing pressure {psia}

= reservoir thickness {ft}

= reservoir drainage area per well {acre}

= wellbore radius {ft}

= Detiz Shape Factor

= total skin factor

87

= viscosity of reservoir gas {cp}

= viscosity of reservoir oil {cp}

= oil formation volume factor {RB/STB}

= gas formation volume factor {RB/SCF}

= solution gas-oil ratio {SCF/STB}

= volatilized oil-gas ratio {STB/SCF}

The first terms in the last bracket of Equation 4-132 and Equation 4-133 represent fluid

production that comes from the reservoir oil, while the second terms represent fluid production

that comes from the reservoir gas. However, for a gas condensate reservoir, reservoir oil is

typically assumed to be immobile ( ). Thus, gas flow rate ( ) and oil flow rate ( )

equations are simplified into Equation 4-134 and Equation 4-135, respectively

Equation 4-134

Equation 4-135

88

In Equation 4-134 and Equation 4-135, only gas and oil flow rates ( and ) or

bottomhole flowing pressure ( ) could be specified. Absolute reservoir permeability ( ),

reservoir thickness ( ), drainage area per well ( ), wellbore radius ( ), and Deitz shape factor

( ) are required reservoir data. Gas formation volume factor ( ), volatilized oil-gas ratio ( ),

and specific gravity of gas ( ) are functions of average reservoir pressure ( ). Average

reservoir pressure ( ) is obtained from the SGCT (zero-dimensional material balance)

subroutine. Gas viscosity ( ) is calculated from correlation proposed by Lee et al in 1966

presented as Equation 4-24 through Equation 4-27.

Even though the relative permeability to oil ( ) can be safely assumed to remain zero

or close to zero during depletion of a gas condensate reservoir, the relative permeability to gas is

not expected to remain equal to one in the presence of condensate. Typically, the mobility of the

gas phase and thus its relative permeability are expected to decrease with increased condensate

drop out. The relative permeability of the gas phase will be further hindered if average water

saturation in the reservoir increases because of the presence of an active water drive. Relative

permeability of reservoir gas ( ) is a function of the average reservoir gas saturation (

). In this study, there are two input options for relative permeability data. The first option

is to input gas saturations and their corresponding values manually in a tabular form. Such

data could be obtained from core study results performed in a laboratory. The second option is to

use a correlation for three-phase relative permeability. Any relative permeability model can be

utilized; however, Naar, Henderson, and Wygal’s model (Ertekin et al, 2001) has been used as a

default model in this work. The correlation is shown in Equation 4-136.

Equation 4-136

89

where:

= average reservoir gas saturation

= connate water saturation

= average reservoir gas saturation

= coefficient to adjust relative permeability of reservoir gas, typically

one when is expected to take value of one at the end point ( = 1-

)

Gas saturation ( ) can change from a minimum gas saturation ( ) to a maximum gas

saturation ( ). Relative permeability of reservoir gas ( ) is equal to zero at the minimum

gas saturation ( ) or lower. Connate water saturation ( ) is a required reservoir data. The

coefficient, , is used to adjust relative permeability anchor point at the initial gas saturation

( ). If at the initial gas saturation, or if there is no further core or lab data

available indicating otherwise, adjustment coefficient value is set to 1.0.

Skin factor is dimensionless pressure drop around the wellbore, which accounts for the

differences between reservoir model’s analytical assumptions and actual conditions in reservoir

flow. Total skin factor ( ) consists of mechanical skin ( ) and non-Darcy skin ( ).

Mechanical skin can be estimated from pressure transient analysis or from other analogous

approaches such as type curve matching. Non-Darcy coefficient ( ) can be obtained from the

analysis of multi-rate well test, or analogous approaches, or from Equation 4-138 (Lee et al,

2003. Eq. 3.19) if its required input data is known

90

Equation 4-137

Equation 4-138

where:

= mechanical skin factor

= non-Darcy coefficient {Day/SCF}

= flow rate of surface gas {SCF/D}

= turbulence parameter

= effective permeability of reservoir gas ( ) {md}

= molecular weight of reservoir gas {lbm/lbmol}

= pressure at standard condition {14.7 psia}

= reservoir thickness {ft}

= wellbore radius {ft}

= temperature at standard condition {520 R}

= viscosity of reservoir gas at bottomhole flowing pressure {cp}

Total skin factor ( ) is a function of gas flow rate, because of the presence of the non-

Darcy skin component, while gas flow rate also is function of skin factor. Thus, Equation 4-134

cannot be solved explicitly for gas flow rate. When the non-Darcy component is expected to be

significant, Equation 4-134 can be recast in terms of a quadratic expression in gas flow rate ( )

91

which can be solved analytically. Alternatively, Equation 4-134 can be solved directly by

implementing an iterative numerical approach. The latter is the approach employed in this study.

4.4.2 Tubing Performance Relationships

For a gas condensate reservoir, condensate drop out is not only expected to occur at

reservoir conditions but also along the surface depletion path as produced fluids make their way

to the surface. For the case of the two-phase two-pseudo component model, the amount of

reservoir condensation can be estimated for the isothermal reservoir conditions using the concept

of volatilized oil-gas ratio. The amount of condensate at the surface is estimated using the surface

pseudo-component concept. However, the table of ―black oil‖ standard PVT properties provides

no information about how much condensate can be expected as a function of both pressure and

temperature changes inside the well tubing during the wellstream fluid travel from reservoir to

surface conditions. The difference in the values of volatilized oil-gas ratio between for any two

points of pressure inside the tubing would provide a measure of gas condensation—but assuming

that those two points are found at separator temperature conditions. To overcome this problem,

this study invokes the homogeneous flow assumption at tubing conditions and applies the well-

known expressions for the flow of gases in a pipeline, adjusted according to an appropriate value

of tubing efficiency. The homogeneous single phase flow equation used to calculate gas flow rate

( ) for a given downstream pressure ( ), and vice versa is shown in Equation 4-139. Pipeline

efficiency factor ( ) is defined to account for the extra pressure drop that should be

expected due to presence of liquid phase.

92

Equation 4-139

Equation 4-140

where:

= gas flow rate {SCF/D}

= upstream pressure {psia}

= downstream pressure {psia}

= pressure at standard condition {14.7 psia}

= pipe section average temperature {R}

= temperature at standard condition {520 R}

= average compressibility factor

= Fanning’s fraction factor

= specific gravity of gas

= tubing diameter {inch}

= tubing length {mile}

= difference in elevation of downstream and upstream {ft}

= efficiency factor of tubing

93

Upstream pressure ( ), tubing diameter ( ), tubing length ( ), elevation at upstream

( ) and downstream ( ) of the tubing are required input data. Difference in elevation is

calculated as . Tubing temperature changes are assumed to follow the geothermal

gradient. An average temperature ( ) is assumed to be equal to an average between

temperatures at upstream ( ) and downstream ( ) nodes for any given tubing section. Average

pressure ( ) is calculated from Equation 4-141. Average compressibility factor ( ) is

calculated at average pressure ( ) and average temperature ( ).

Equation 4-141

In Equation 4-140 above, the Fanning’s friction factor ( ) is equal to a quarter of

Moody’s friction factor ( ). The Moody’s fraction factor ( ) can be calculated from Equation

4-142 (Colebrook, 1939). This equation is solved using iterative procedure because Moody’s

friction factor appears implicitly on both sides of the equation. The Reynolds number is

calculated from Equation 4-143. In situ fluid density ( ) and velocity ( ) of gas are evaluated at

average pressure and temperature using real gas equation of state. Gas viscosity is calculated

from the correlation proposed by Lee et al in 1966.

Equation 4-142

94

Equation 4-143

where:

= Moody’s friction factor

= tubing diameter {inch}

= tubing roughness {ft}

= Reynolds number

= fluid density {lbm/ft3}

= fluid velocity {ft/sec}

= fluid viscosity {cp}

The concept of tubing efficiency factor ( ) is applied to for the detrimental

presence of a liquid phase on pipe performance. For gas condensate fluid, amount of liquid phase

in pipeline is partially related the change in volatilized oil-gas ratio ( ) and tubing flowing

pressure, which both of them are also functions of reservoir pressure. Therefore, the efficiency

factor would also implicitly depend on reservoir pressure. In this study, a single value of tubing

efficiency factor, assumed to be representative of the entire tubing flow, is employed. The

efficiency factor should be calculated in such a way that the calculated flow rate and pressure

would match actual production data for the period of interest.

During calculations, the tubing is divided into several pipe sections. The calculation is

performed section by section, starting at the bottommost section and marching towards the

topmost section of the tubing. The upstream pressure of the bottommost section is the bottomhole

flowing pressure, while the upstream pressure of the upper section is the calculated downstream

95

pressure from the lower section. Wellhead pressure represents the downstream pressure of the

topmost section of the tubing.

TPR calculations can be performed in two ways. One possibility is calculating wellhead

pressure ( ) as a function of gas flow rate ( ). For such scenario, downstream pressure ( )

in gas flow equation is sequentially solved starting at the bottommost section until reaching the

topmost section of the tubing. For each section, an iterative process is needed to determine

downstream pressure ( ) because downstream pressure ( ) affects the value of average

compressibility factor ( ) and Reynolds number ( ). The downstream pressure ( ) of the

topmost section of the tubing is the desired target.

The second possibility for TPR calculations is finding the corresponding gas flow rate

( ) for a given wellhead pressure ( ). In this case, initial guess of gas flow rate ( ) is

calculated using well-known Weymouth correlation. Then, wellhead pressures ( ) are

calculated for the initial guess of gas flow rate ( ). Based on the ensuing sensitivity analysis,

incremental wellhead pressure ( ) and incremental gas flow rate ( ) can be estimated.

The gas flow rate ( ) is now updated based on the derivative of gas flow rate ( )

and the difference between calculated and specified wellhead pressures ( ). This procedure is

repeated until the difference between calculated and specified wellhead pressure is less than a

prescribed tolerance value.

96

4.4.3 Nodal Analysis

In petroleum engineering, nodal analysis is used to determine pressure and flow rate at

some node or location of interest in the production system. In this study, sand-face location is the

selected node or location of interest. Incoming gas flow rate to the node (―inflow‖) can be

calculated from inflow performance relationships (IPR), while outgoing gas rate from the node

(―outflow‖) can be calculated using tubing performance relationships (TPR). Nodal analysis is

then performed to determine pressure ( ) and flow rate ( ) that satisfies both IPR and TPR

relations at the same time – which correspond to the point where ―inflow‖ and ―outflow‖ curve

cross each other, as depicted in Figure 4-5.

Figure 4-5: Graphical Representation of Nodal Analysis

Figure 4-5 shows the graphical representation of nodal analysis method. IPR curve is

constructed using Equation 4-134, and TPR curve is constructed using Equation 4-139. The

Pw

f-

Bo

tto

mh

ole

Flo

win

g P

ress

ure

qgsc - Gas Flow Rate

IPR TPR

Tubing Performance Relationship

Inflow Performance Relationship

Pwf and qgsc

that satisfy both IPR & TPR

97

intersection between IPR and TPR curves represents the solution of the nodal analysis problem.

As it can be seen from the figure, as the bottomhole flowing pressure ( ) is lower than the

intersection pressure, more gas would be able to flow from reservoir into the wellbore; however,

the tubing would not be able to deliver all of gas to the surface for the given pressure drop. In

contrast, if the bottomhole flowing pressure ( ) is higher than the intersection pressure, less

gas is able to flow from reservoir into the wellbore even though the tubing would be able to

deliver more than that. Both scenarios are not physically possible for a steady state condition

where the flow from the reservoir and in the tubing must be equal to each other. There is only one

possible flow condition which is found at the intersection point. At the intersection pressure, all

of gas which flows from the reservoir into the wellbore can be delivered to the surface production

system.

The solution of the IPR and TRP point of intersection requires an iterative procedure. In

this study, the bi-section method is applied. Reservoir pressure ( ) and wellhead pressure ( )

are the independent variables while bottomhole flowing pressure ( ) and gas flow rate ( )

are the dependent variables. Initial guess of bottomhole flowing pressure ( ) is taken as an

average between the higher pressure boundary which is equal to reservoir pressure ( ) and lower

pressure boundary which is equal to wellhead pressure ( ) for the first iteration. Then, gas flow

rates ( ) from IPR and TPR correlations are calculated for the current guess of bottomhole

flowing pressure ( ). If gas flow rate calculated from the IPR curve is less than gas flow rate

from TPR, then the higher pressure boundary is replaced by current bottomhole flowing pressure

( ). If not, the lower pressure boundary is replaced by current bottomhole flowing pressure

( ) instead. New iteration starts by averaging the bottomhole flowing pressure ( ) from the

updated higher and lower pressure boundaries. The iterative process is repeated until the

98

difference between calculated gas flow rate ( ) from IPR and TPR is less than a prescribed

tolerance.

99

4.5 Field Performance Prediction

Field performance predictions are based on the integration of reservoir tank models with

flow rates and flowing pressure (IPR/TRP) models, subjected to a given set of production

constraints, with the goal of predicting pressures (reservoir, wellhead, bottomhole), flow rate, and

cumulative production evolution with respect to reservoir production time. The production time

variable, which is eliminated during the development of reservoir tank or zero-dimensional

models, is now calculated by combining tank models and flow rate models (IPR/TRP) in the same

calculation.

In this work, substantial drilling capacity is assumed; thus build-up period is neglected.

Field life is divided into two periods: the plateau period and decline period. All wells are put on

production at the beginning of plateau period. During the plateau period, gas flow rate ( ) is

kept constant by adjusting wellhead pressure ( ). During the decline period, wellhead pressure

( ) is kept constant at the minimum allowable wellhead pressure. The field is abandoned when

the reservoir pressure reaches the specified abandonment pressure. The following sections

described the algorithm for predicting field performance in a step-by-step procedure.

100

4.5.1 Performance during Plateau Period

During plateau period, the delivered gas flow rate ( ) from the field is kept constant.

In order to achieve this, in spite of decreasing pressure and gas flow capacity in the reservoir,

wellhead pressures ( ) are adjusted throughout the plateau period in order to honor the

constant production specification. Naturally, the reservoir must have the capacity (in terms of

reserves or OGIP) of delivering such target plateau rate throughout the plateau period. Therefore,

the gas flow rate during plateau period (

) is constrained by the feasibility of productive

reservoir performance at the end of plateau period. The following procedure is used to predict the

field performance during plateau period.

Estimation of Cumulative Production at End of Plateau ( )

Cumulative production at end of plateau is calculated from target recovery factor at end

of plateau ( ) (which is considered a given or specified variable provided to the model)

times original gas in place ( ) (which must be known before field performance predictions can be

undertaken).

Estimation of Reservoir Pressure at End of Plateau (

)

Reservoir pressure at end of plateau (

) is calculated from cumulative gas

production at end of plateau ( ) and the SGCT model. In the SGCT model, reservoir

pressure ( ) is an independent variable, and cumulative gas recovery ( ) is the dependent

variable. Cumulative gas production ( ) is calculated from cumulative gas recovery ( )

times original gas in place ( ). Therefore, finding reservoir pressure ( ) at a given cumulative

101

gas production ( ) requires an iterative process or pressure that is build around matching the

prescribed gas production at the end of plateau ( ). Bi-section iterative method is utilized

in this study.

Estimation of Plateau Gas Flow Rate (

)

Plateau gas flow rate (

) is calculated using the nodal analysis method. Because in

this calculation the minimum allowable wellhead pressure ( ) fixes the gas rate at the end of

plateau, nodal analysis is applied at the prevailing conditions at the end of the plateau period. For

the IPR calculation, for example, the reservoir pressure ( ) is equal to reservoir pressure at end

of plateau (

). For the TPR calculation, wellhead pressure ( ) is fixed to be equal to

minimum allowable wellhead pressure ( ). Intersection of IPR and TPR curves are the plateau

gas flow rate per well (

) and bottomhole flowing pressure at end of plateau (

).

Total plateau gas flow rate (

) is equal to the plateau rate per well (

) times total

number of wells.

Estimation of Reservoir Pressure ( ), Bottomhole Flowing Pressure ( ), and Wellhead

Pressure ( ) during plateau period

During the plateau period, reservoir pressure ( ) decreases continuously from initial

reservoir pressure ( ) to reservoir pressure at end of plateau (

) based on the

predictions from the SGCT model. Bottomhole flowing pressures ( ) are calculated as a

function of plateau gas flow rate per well (

) at different reservoir pressures ( ) by

implementing the IPR equation (Equation 4-134). Wellhead pressures ( ) are calculated as a

102

function of plateau gas flow rate per well (

) at different bottomhole flowing pressures

( ) by implementing the TPR calculation.

Estimation of Gas Flow Rate ( ) and Oil Flow Rate ( )

Total gas flow rate production ( ) during plateau period is constant and equal to the

prescribed plateau gas flow rate (

) for the entire field. Total oil flow rates ( ) are

calculated from total gas flow rate ( ) times volatilized oil-gas ratios ( ). Volatilized oil-gas

ratios ( ) at different reservoir pressures ( ) are linearly interpolated from the table of standard

PVT properties.

Estimation of Cumulative Gas Production ( ) and Cumulative Oil Production ( )

Cumulative gas production ( ) is calculated as cumulative gas recovery ( ) times

original gas in place ( ). Cumulative oil production ( ) is calculated as cumulative oil

recoveries ( ) times original oil in place ( ). Cumulative gas ( ) and cumulative oil

( ) recoveries at different reservoir pressures ( ) are calculated from the SGCT model.

Estimation of Production Time ( )

Production times ( ) at different reservoir pressures during the plateau period are

calculated from cumulative gas productions ( ) divided by total plateau gas flow rate

(

). Plateau time ( ) is equal to cumulative gas production at end of plateau

( ) over total plateau gas flow rate (

).

103

4.5.2 Performance during Decline Period

During decline period, wellhead pressure ( ) is kept at a constant value equal to the

minimum allowable wellhead pressure ( ). As a consequence of this, total gas flow rate ( )

would continuously decrease as a function of reservoir pressure ( ). The following procedure is

used to predict field performance during decline period.

Estimation of Reservoir Pressure ( ), Bottomhole Flowing Pressure ( ), and Wellhead

Pressure ( ) during decline

During the decline period, reservoir pressure ( ) continues to decreases from reservoir

pressure at the end of plateau (

) to abandonment reservoir pressure ( ) based on

predictions from the SGCT model. Bottomhole flowing pressures ( ) are calculated using

nodal analysis. In the IPR calculation, reservoir pressures ( ) change with production. Wellhead

pressure is fixed ( ) in the TPR calculation. Intersections of IPR and TPR curves at different

reservoir pressures ( ) yield the resulting bottomhole flowing pressures ( ) and gas flow rates

per well ( ) for the decline period.

Estimation of Gas Flow Rate ( ) and Oil Flow Rate ( )

Gas flow rates per well ( ) at different reservoir pressures ( ) are calculated along

with bottomhole flowing pressures ( ) using the nodal analysis discussed earlier. Total gas

flow rate ( ) is equal to gas flow rate per well ( ) times total number of wells. Oil flow

rate ( ) is calculated from total gas flow rate ( ) times the volatilized oil-gas ratio

104

( ).Volatilized oil-gas ratios ( ) at different reservoir pressures ( ) are linearly interpolated

from the table of standard PVT properties.

Estimation of Cumulative Gas Production ( ) and Cumulative Oil Production ( )

Cumulative gas productions ( ) and cumulative oil production ( ) calculation for

decline period follows the same protocol discussed for the plateau period.

Estimation of Decline Rate ( i)

If the reservoir decline behavior is assumed to be an exponential decline, the decline rate

( ) is expected to remain constant throughout decline period. Overall decline rate ( ) is

calculated from difference between total gas flow rates at end of plateau (

) and

abandonment ( ) over difference between cumulative gas production at end of plateau

( ) and abandonment ( ).

Equation 4-144

Estimation of Production Time ( )

Production times ( ) at different reservoir pressures ( ) are calculated from Equation

4-145. The calculation is based on the exponential decline assumption and constant decline rate

105

( ) discussed earlier. Abandonment time ( ) is calculated from Equation 4-145 at for the

condition at which flow rate equals abandonment total gas flow rate ( ).

Equation 4-145

106

4.5.3 Annual Production Calculation

Annual hydrocarbon production is the principal input for economic evaluation. Annual

production volumes during plateau period can be straightforwardly calculated because rates are

constant for the case of gas production. For the decline period, however, calculating annual

production is more involved because flow rates are time dependent. Iterative procedures can be

used to find exact values of annual production during decline period; however, because of the

exponential decline assumption, produced volumes can be determined directly, as outlined below

for annual gas and oil production calculations. Annual production time ( ) is defined as Equation

4-146. It converts production time ( ) from ―day‖ unit to ―year‖ unit. Annual production time ( )

begin from zero at initial condition.

Equation 4-146

Annual Gas Production

Annual gas production at year ( ) is calculated from difference between cumulative

gas production at year ( ) and cumulative gas production at year ( ).

Equation 4-147

107

For plateau period ( ), cumulative gas production at year ( ) is calculated

from Equation 4-148. For decline period ( ), cumulative gas production at

year ( ) is calculated from Equation 4-149.

Equation 4-148

Equation 4-149

Annual Oil Production

Annual oil production at year ( ) is calculated from the difference between

cumulative oil production at year ( ) and cumulative oil production at year ( ).

Cumulative oil production ( ) is calculated using linear interpolation with respect to cumulative

gas production ( ).

Equation 4-150

Instantaneous and Annual Average Flow Rates

Instantaneous flow rates of gas ( ) and oil ( ) are calculated using linear

interpolation with respect to cumulative gas production ( ). Annual average flow rates of gas

( ) and oil ( ) are calculated from the annual gas production ( ) and the annual oil

108

production ( ) divided by 365.25, respectively. In general, annual average rates can be used as

close approximation of instantaneous rates. They can also be converted to annual production

volumes easily. Because of this, annual average rates are more meaningful to report than

instantaneous flow rates for the purpose of economic analysis.

109

4.6 Economic Analysis and Field Optimization

Economic analysis is the analytical method that quantifies economic performance or

monetary value of a field investment project and provides a meaningful metric for the

optimization of field operations. Economic model which used in this study is based on typical

cash flow before tax regime (Mian, 2002).

An economic evaluation subroutine has been developed to perform the economic analysis

in this study. It consists of three main parts: compilation of production data, calculation of net

present value (NPV), and calculation of rate of return (ROR). The economic subroutine has been

constructed independently of the subroutine for field performance prediction. In this way,

economic analysis can be performed either with or without calculating new field performance

data. This is especially useful when performing parametric studies on economic parameters rather

than field operational constraints.

An optimization subroutine was also developed to find the recommended target recovery

factor at end of plateau to be contracted and total recommended number of wells to be drilled for

optimal field development. The optimization module requires the use of a variety of values of

target recovery factor and total number of wells that need to be economically screened. The

subroutine would repeatedly call the field performance prediction subroutine for a number of

different values of target plateau recovery factor and total number of wells. Economic analysis is

then performed for each development scenario and calculated NPVs are stored into an

optimization table or matrix. Production profiles corresponding to each of the investigated target

recovery factor and total numbers of well combinations are stored in the same worksheet.

A separate subroutine is available for the formulation of economic analysis decoupled

from field performance calculations. This subroutine would recall the stored production profiles

which were generated earlier during the economic evaluation discussed above. This subroutine is

110

useful because the generation of production profiles is the step in the analysis that requiring the

most significant computational time, and its availability allows to re-evaluate the sensitivity of

field optimization to different values of economic parameter(s). The subroutine is especially

useful when there are no changes in field operational constraints.

4.6.1 Simplified Economic Model

Net annual productions are calculated from annual production volumes times the net

hydrocarbon interest (Equation 4-151). Annual production volumes can be obtained from the

annual production calculation outlined in Section 4.5.3. The net hydrocarbon fraction or interest

is the fraction of the hydrocarbon production which is earned by the operator or investor. The

remaining portion of hydrocarbon belongs to the owner of land or the mineral rights.

Equation 4-151

Annual revenues are calculated by multiplying the annual production volume with the

estimated price of the product (Equation 4-152). The estimated price of gas has to be provided in

the units of $ per MSCF while the price of oil or condensate has to be provided in the unit of $

per STB. Generally, gas price is quoted in the unit of $ per MMBTU, which can be converted to $

per MSCF by multiplying $/MMBTU times the gas heating value {BTU/SCF} and dividing

through by 1000. Total annual revenue becomes the sum of annual revenue from gas and oil

(Equation 4-153).

111

Equation 4-152

Equation 4-153

Capital expenditure (CAPEX) is a group of one-time costs which occur in order to make

production possible. Total CAPEX consists of fixed CAPEX such as platform costs, flowline

costs, and production facilities costs, and variable CAPEX includes the variable costs of

operations such as drilling and completion (D&C). D&C cost is equal to D&C cost per well times

total number of wells (Equation 4-155). For this simplified economic model, CAPEX is assumed

to occur at the beginning of the project.

Equation 4-154

Equation 4-155

Operating expenditure (OPEX) is a group of costs which occur periodically in order to

maintain the day-to-day operation. OPEX may include maintenance cost, utilities cost, overhead

cost, production cost, etc. In this work, OPEX has to be given in the unit of $ per month. Annual

OPEX is calculated from monthly OPEX times 12 months (Equation 4-156).

112

Equation 4-156

Severance tax is a government or state tax which is imposed on the production of non-

renewable resources such as oil and natural gas. Ad Valorem tax is a tax which imposed at the

time of transaction. In this economic model, severance taxes for oil and gas have to be given in

terms of a percentage of the total oil and gas revenue, respectively. Annual severance taxes are

calculated from annual revenues times the severance tax rate (Equation 4-157). Ad Valorem tax

has to be given in percentage of total revenue. Annual Ad Valorem tax is calculated from annual

total revenue times the Ad Valorem tax rate (Equation 4-158). Annual total tax is summation of

severance taxes and Ad Valorem tax (Equation 4-159).

Equation 4-157

Equation 4-158

Equation 4-159

113

Annual expenditure is the sum of total CAPEX, annual OPEX, and annual total tax

(Equation 4-160). Annual net cash flow is the difference between annual total revenue and annual

expenditure (Equation 4-161). Cumulative net cash flow at year is the summation of annual net

cash flow from the beginning ( ) to year .

Equation 4-160

Equation 4-161

Equation 4-162

Commodity prices and operating cost can be escalated according to the economic

inflation. In this simplified economic model, gas price, oil price, and monthly OPEX are escalated

independently. The escalation is applied from the start to the end of the project. Today’s product

prices, monthly OPEX, and their escalation rates have to be provided to the model. Future prices

and monthly OPEX are calculated from Equation 4-163 and Equation 4-164. Un-escalated

economic analysis, which is required for some official reports, can be evaluated by specifying all

escalation rates to be equal to zero.

114

Equation 4-163

Equation 4-164

Annual net cash flow in the future is discounted to today’s equivalent value using concept

of ―time value of money‖. Annual discounted net cash flow at year is calculated using Equation

4-165. Annual net cash flow at year can be obtained from Equation 4-161. The effective interest

rate, which is interest rate applied on an annual basis, has to be provided. Cumulative discounted

net cash flow at year is the summation of annual discounted net cash flow from the beginning

( ) to year .

Equation 4-165

Equation 4-166

115

Net present value ( ) is equal to cumulative discounted net cash flow at abandonment.

The rate of return ( ) is the interest rate which results in zero NPV. The NPV profile is

calculated from evaluating NPV for a variety of interest rates. In this model, the interest rates are

varied from 5% to 40%. However, the range can be adjusted depending on the analyst and desired

economic results.

116

4.6.2 Field Optimization

In this study, target recovery factor at end of plateau and total number of wells are targets

for optimization because they can be controlled by the operator of the field. Net present values

( ) at differences target recovery factor and total number of wells are evaluated and

compared. The combination of target recovery factor and total number of wells which providing

the global maximum NPV is the desired result.

Figure 4-6: Graphical Representation of Field Optimization

Production profiles of oil and gas at differences target recovery factor and total number of

wells are calculated based on field performance prediction concept (Section 4.5). The predicted

results are then imported into the simplified economic model (Section 4.6.1). The production

scenario which provides the highest NPV is selected.

TotalNumberof Wells

Net

Pre

sen

t V

alu

e

Target Recovery Factor at End of Plateau

Optimum NPV

Chapter 5

Model Performance

5.1 Simulation of Standard PVT “Black Oil” Properties

A model has been developed to simulate standard PVT properties from a phase behavior

protocol. Input data presented in Appendix A is set to the subroutine and standard ―black oil‖

PVT properties - gas formation volume factor ( ), oil formation volume factor ( ), volatilized

oil-gas ratio ( ), and solution gas-oil ratio ( ) - are calculated based on fluid compositional

data. Section 5.1.1 shows calculation results from the procedure described in Section 4.2.3.

Section 5.1.2 discusses the limitations inherent to representing a multi-component hydrocarbon

mixture as a binary pseudo-component fluid model. Section 5.1.3 elaborates about the impact of

such limitations on the values and behavior of standard PVT properties.

5.1.1 Simulated Standard PVT Properties

The standard PVT properties calculated from data set in Appendix A are shown in Figure

5-1 and Figure 5-2. The values of these properties depend on pressure (reservoir and surface),

temperature (reservoir and surface), and original composition of in-situ reservoir fluid. The trends

depicted in Figure 5-1 and Figure 5-2 are consistent with the typical phase behavior of gas

condensate reservoir fluids discussed in section 4.2.1 and the standard PVT properties for the

Anschutz Ranch East rich-gas condensate presented by Walsh and Lake (2003).

118

Figure 5-1: Simulated Gas Formation Volume Factor

and Volatilized Oil-Gas Ratio of Gas Condensate

For this reservoir fluid, gas formation volume factor ( ) equals 1.040 RB/MSCF at the

initial reservoir pressure of 4000 psia. As shown in Figure 5-1, its value increases with decreasing

reservoir pressure with an increased slope at lower reservoir pressures. At final abandonment

pressure of 500 psia, gas formation volume factor ( ) reached the value of 7.816 RB/MSCF.

Volatilized oil-gas ratio ( ) remains initially constant at the value of 205.9 STB/MMSCF for

reservoir pressures higher than the fluid’s dew point. Its value starts decreasing as soon as the

pressure decreases below dew point conditions. The decreasing slope of volatilized oil-gas ratio

( ) is largest at conditions around dew point conditions and lower as reservoir pressure

decreases. As the pressure continues to decrease, volatilized oil-gas ratio ( ) starts increasing

again because of re-vaporization of oil. Volatilized oil-gas ratio ( ) reaches a minimum value of

63.5 STB/MMSCF around 1100 psia before increasing up to 80.8 STB/MMSCF at the final

pressure of 500 psia.

0

50

100

150

200

250

0

2

4

6

8

10

0 1000 2000 3000 4000 5000

Rv

-V

ola

tiliz

ed

Oil

-Gas

Rat

io (

STB

/MM

SCF)

Bg

-G

as F

orm

atio

n V

olu

me

Fac

tor

(RB

/MSC

F)

Reservoir Pressure (psia)

Bg (RB/MSCF) Rv (STB/MMSCF)

Dew Point Pressure

Rv

Bg

119

Figure 5-2: Simulated Oil Formation Volume Factor

and Solution Gas-Oil Ratio of Gas Condensate

At reservoir pressure higher than dew point pressure, oil formation volume factor ( )

and solution gas-oil ratio ( ) are not actually defined because there is no free liquid phase in the

reservoir. However, oil formation volume factor ( ) and solution gas-oil ratio ( ) values above

the dew point can be calculated from Equation 4-61 and Equation 4-62. Those relationships are

defined force the Material Balance Equation (MBE) for gas condensate to collapse to the MBE

for wet gas at pressures above the dew point. As shown in Figure 5-2, oil formation volume factor

( ) starts at 2.858 RB/STB around dew point conditions and monotonically decreases to a value

of 1.206 RB/STB at the final pressure of 500 psia. Similarly, solution gas-oil ratio ( ) begins at

2067 SCF/STB around dew point conditions and then monotonically decreases to a value of 61.5

SCF/STB at the final pressure.

0

1000

2000

3000

4000

5000

6000

7000

0

1

2

3

4

5

6

7

0 1000 2000 3000 4000 5000

Rs-

Solu

tio

n G

as-O

il R

atio

(SC

F/ST

B)

Bo

-O

il Fo

rmat

ion

Vo

lum

e F

acto

r (R

B/S

TB)

Reservoir Pressure (psia)

Bo (RB/STB) Rs (SCF/STB)

Dew Point Pressure

Bo

Rs

120

Figure 5-3: Simulated Specific Gravity of Reservoir Gas

In addition, specific gravity ( ) of reservoir gas is also estimated. The results of this

estimation are shown in Figure 5-3. Above the dew point, specify gravity ( ) of reservoir gas is

constant and equal to 1.147. After that, its value decreases with decreasing reservoir pressure to

the minimum value of 0.876 at 1150 psia and then slightly increases back to 0.919 at the final

pressure of 500 psia due to condensate re-vaporization.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 500 1000 1500 2000 2500 3000 3500 4000 4500

SG -

Spe

cifi

c G

ravi

ty o

f R

ese

rvo

ir G

as

Reservoir Pressure (psia)

Specific Gravity of Reservoir Gas

Dew Point Pressure

SG

121

5.1.2 Limitations of Pseudo Component Model

The calculated volumes of surface gas and stock-tank oil pseudo components in reservoir

gas, reservoir oil, and cumulative gas production below dew point pressure are displayed in

Figure 5-4 and Figure 5-5, as they were estimated during the simulation of a CVD experiment

using a phase behavior model. They are calculated from 1000 MSCF of equivalent gas at dew

point pressure.

Figure 5-4: Volumes of Surface Gas Pseudo Component in

Reservoir Gas Reservoir Oil, and Cumulative Gas Production

As shown in Figure 5-4, the amount of surface gas pseudo component remaining in the

reservoir gas phase ( ) decreases from a value of 847 MSCF at the dew point to a value of 121

MSCF at the final pressure because of continuous gas production and retrograde condensation of

reservoir gas. Amounts of surface gas pseudo component in the reservoir oil phase ( ) starts at

0

100

200

300

400

500

600

700

800

900

1000

0 500 1000 1500 2000 2500 3000 3500

G -

Surf

ace

Gas

Pse

ud

o C

om

po

ne

nt

(MSC

F)

Reservoir Pressure (psia)

Gfg (MSCF) Gfo (MSCF) Gp (MSCF) Gfg +Gfo + Gp (MSCF)

Gfg

Gp

Gfg+Gfo+Gp

GfoDew Point

Pressure

847

121

735

5

861

122

zero at dew point pressure because there is no reservoir oil. Then, it increases to the maximum

volume of 92 MSCF at 2700 psia before decreasing to 5 MSCF at 500 psia. This reversing trend

is dominated by the changing amounts of reservoir oil volume or oil saturation during depletion.

Cumulative gas production ( ), which is the volume of surface gas pseudo component

recovered from the production of excess gas out of the PVT cell, increases from zero at dew point

pressure to 735 MSCF at the final pressure.

Figure 5-5: Volumes of Stock-Tank Oil Pseudo Component in

Reservoir Gas, Reservoir Oil, and Cumulative Oil Production

Figure 5-5 shows that the amount of stock-tank oil pseudo-component in the reservoir gas

phase ( ) decreases from a value of 174 STB at dew point pressure to a value of 10 STB at the

final pressure point due to gas production and retrograde condensation of reservoir gas. The

amount of stock-tank oil pseudo-component in reservoir oil ( ) increases from zero at dew

point pressure because of the lack of a reservoir oil phase and reaches a maximum volume of 96

0

20

40

60

80

100

120

140

160

180

200

0 500 1000 1500 2000 2500 3000 3500

N -

Sto

ck-T

ank

Oil

Pse

ud

o C

om

po

ne

nt

(STB

)

Reservoir Pressure (psia)

Nfg (STB) Nfo (STB) Np (STB) Nfg +Nfo + Np (STB)

Nfo

Nfg

Nfg+Nfo+Np

Np

DewPoint

Pressure

174

10

88

67

165

123

STB at 1450 psia before decreasing to 88 STB at the final pressure. This reversing trend mainly

stems from changes in reservoir oil saturation in the reservoir during depletion. Cumulative oil

production ( ), which is the volume of stock-tank oil pseudo-component recovered from the

excess gas produced from the PVT cell, increases from zero at dew point pressure to 67 STB at

the final pressure.

Let us now consider how the basic statements of species conservation are being honored

during the simulated depletion. The sum of the amounts of surface gas pseudo-component in

reservoir gas, in reservoir oil, and cumulative gas production ( ) in Figure 5-4

should be always equal to the total amount of surface gas pseudo component originally present in

reservoir gas at dew point conditions if the surface gas pseudo-component was to be fully

conserved throughout the simulated experiment. However, Figure 5-4 clearly shows that the total

amount of surface gas pseudo component ( ) is not constant throughout the

simulated process; instead, it increases from 847 MSCF at the dew point to 861 MSCF at the final

pressure. This suggests that total mass of surface gas pseudo component actually increases with

decreasing reservoir pressure which is physically impossible as is in violation of mass

conservation. Similarly, the fact that the total amount of stock-tank oil pseudo component

( ) decreases from 174 STB at dew point conditions to 165 STB at the final

pressure, suggests that total mass of stock-tank oil pseudo-component actually decreases with

decreasing reservoir pressure. These unphysical trends are caused by one of the key assumptions

used in pseudo component model, which establishes that the properties of the surface gas and

stock-tank oil pseudo-components are supposed equal and unchanging during depletion.

In reality, the properties of pseudo components, which are actually two multi-component

mixtures in their own right, do change throughout reservoir depletion. Compositions of reservoir

gas and reservoir oil change continuously because of retrograde condensation. In addition, there

are three points of surface separation; first stage separator, second stage separator, and stock tank.

124

Figure 5-6 displays the estimated densities of surface gas and stock-tank oil pseudo components,

as calculated by the Phase Behavior Model. They clearly suggest that fluid properties of these

pseudo-components are not necessarily constant, although they do not change dramatically during

depletion.

Figure 5-6: Densities of Surface Gas and Stock-Tank Oil Pseudo Components

at First Stage Separator, Second Stage Separator and Stock Tank Condition

Figure 5-7 shows the amount of stock-tank oil pseudo-component in the reservoir gas

phase ( ), reservoir oil phase ( ) and cumulative production ( ) in term of gas-equivalent

volume. The conversion from oil volume to equivalent-gas volume is based on molar

equivalency. Gas-oil equivalency factor ( ) is calculated from Equation 5-1. Densities and

molecular weights of stock-tank oil from reservoir gas, reservoir oil, and cumulative oil

production are directly obtained from Phase Behavior Model.

0

10

20

30

40

50

0.0

0.6

1.2

1.8

2.4

3.0

0 500 1000 1500 2000 2500 3000 3500

De

nsi

ty o

f St

ock

-Tan

k O

il (l

bm

/ft3

)

De

nsi

ty o

f Su

rfac

e G

as (

lbm

/ft3 )

Reservoir Pressure (psia)

Gfg at Sep 1 (lbm/ft3) Gfg at Sep 2 (lbm/ft3) Gfg at STO (lbm/ft3) Gfo at Sep 1 (lbm/ft3)

Gfo at Sep 2 (lbm/ft3) Gfo at STO (lbm/ft3) Nfg at STO (lbm/ft3) Nfo at STO (lbm/ft3)

Nfg at STO

Nfo at STO

Gfg at Sep1

Gfo at Sep1

Gfg at Sep2

Gfg at STO

Gfo at Sep2

Gfo at STO

125

Equation 5-1

Figure 5-7: Volumes of Stock-Tank Oil Pseudo Component in

Reservoir Gas, Reservoir Oil, and Cumulative Oil Production

in term of Gas-Equivalent

Figure 5-8 shows the volume of surface gas ( ), volume of stock-tank oil

( ) expressed in term of gas-equivalent volume, and total volume from both

surface gas and stock-tank oil volumes. The total volumes are constant and equal to 1000 MSCF

which is the original volume at dew point conditions. These results clearly prove that pseudo

component model is able to honor overall material balance, but not species material balance.

0

20

40

60

80

100

120

140

160

180

0 500 1000 1500 2000 2500 3000 3500

GE

-Su

rfac

e G

as E

qu

ival

en

t (M

SCF)

Reservoir Pressure (psia)

Nfg Nfo Np Nfg+Nfo+Np

Nfg

Ngo

Nfg+Nfo+Np

Np

Dew Point

Pressure

153

62

67

10

139

126

Figure 5-8: Total Volumes of Stock-Tank Oil Pseudo Component and

Surface Gas Pseudo Component in term of Gas-Equivalent

0

100

200

300

400

500

600

700

800

900

1000

0 500 1000 1500 2000 2500 3000 3500

GE

-Su

rfac

e G

as E

qu

ival

en

t (M

SCF)

Reservoir Pressure (psia)

Nfg+Nfo+Np Gfg+Gfo+Gp TOTAL

Gfg+Gfo+Gp

Dew Point

Pressure153

847

139

861

Nfg+Nfo+Gp

TOTAL

127

5.1.3 Impact on Standard PVT Properties

The limitations of representing a multi-component hydrocarbon mixture using a binary

pseudo-component model, as discussed in the preceding section, would definitely have an effect

on the calculation of the standard PVT ―black-oil‖ properties. As discussed below, solution gas-

oil ratios ( ) could become negative at low reservoir pressures; or oil formation volume factors

( ) could become over-estimated.

The Walsh-Towler algorithm applies concept of mass balance around the PVT cell by

calculating amounts of pseudo components in the reservoir oil at any pressure level j ( and

) from the differences between total pseudo-component amounts from the previous pressure

level ( and

) and the summation of pseudo-

component amounts in the reservoir gas phase and cumulative production ( and

).

The combination of this mass balance concept and simulated results from a phase

behavior model can lead to negative values of surface gas pseudo component in oil phase ( )

and over-estimated values of stock-tank pseudo component in the oil phase ( ) when reservoir

depletion is extensive (i.e., at low pressures). For example, at final abandonment pressure of 500

psia and using the data of Figure 5-4, the value of surface gas pseudo component in oil phase

( ), which is equal to 5 MSCF based on a rigorous flash calculation, would be equal to -9

MSCF (847 – 735 – 121 = -9) when this mass balance concept is applied. In Figure 5-5, the value

of stock-tank oil pseudo component ( ), which is equal to 88 STB from a flash calculation,

would be equal to 97 STB (174 – 67 – 10 = 97) using the same mass balance concept. Thus, the

resulting solution gas-oil ratio ( ) calculated from Walsh-Towler algorithm at this condition

would become equal to -96 SCF/STB, while 61.5 SCF/STB is the simulated value from flash

128

calculations as shown in Figure 5-2. Similarly, oil formation volume factor ( ) calculated from

Walsh-Towler algorithm would be equal to 1.082 RB/STB, while 1.206 RB/STB is the simulated

value from flash calculations as shown in Figure 5-2.

129

5.2 Zero-Dimensional Material Balance Calculations for Gas Condensates

A reservoir model has been developed in order to predict cumulative gas production,

cumulative oil production, production gas-oil ratio, and average reservoir gas saturation based on

the procedure described in Material Balance Equation for Gas Condensates and Saturation

Calculation sections discussed above. Input data in Appendix A is used by the model, which

generates its outputs as function of reservoir pressure. Section 5.2.1 discusses the simulation

results and section 5.2.2 discusses the significant pitfalls of misusing the proposed gas condensate

fluid tank model for performance prediction for other near-critical fluid: the volatile oil reservoir.

5.2.1 Simulation Results from Gas Condensate Tank Model

Figure 5-9 plots the results of gas condensate tank model for the reservoir scenario

described in Appendix A. Initial reservoir pressure is 4000 psia, dew point pressure is 3031 psia,

and abandonment pressure is 750 psia. In Figure 5-9, cumulative gas recovery ( ) and

cumulative oil recovery ( ) increase with decreasing reservoir pressure. The recovery slopes

are identical for conditions above dew point pressure; however, the increasing trend or slope of

cumulative oil recovery becomes significantly flatter than that of gas for pressures below the dew

point. At abandonment pressure conditions, in this example, cumulative gas and oil recoveries are

equal to 80.6% and 45.7%, respectively. Gas saturation ( ) remains constant at 79% (

) above the dew point. As reservoir pressure decreases below the dew point, gas saturation

( ) decreases to the minimum value of 66.6% at 2350 psia, and then slightly increases back to

69.5% at the abandonment pressure. Production gas-oil ratio ( ), which is the inverse of

volatilized oil-gas ratio ( ) for a gas condensate system whose reservoir condensate remains

immobile, is constant at 4.857 MSCF/STB above the dew point. As reservoir pressure decreases

130

below dew point conditions, production gas-oil ratio ( ) increases to the maximum value of

15.756 MSCF/STB at 1050 psia, and then slightly decreases back to 14.863 MSCF/STB at the

abandonment pressure.

Figure 5-9: Simulated Production Results of Gas Condensate

using Simplified Gas Condensate Tank Model

Qualitatively, the production trends calculated from the gas condensate zero-dimensional

model fully agree with the typical and expected trends observed in the field and through fully-

dimensional numerical reservoir simulation (Walsh and Lake, 2003). A standard numerical

simulator also suggests, and field experience corroborates, that ultimate gas recovery - i.e.,

cumulative gas recovery at abandonment conditions - is much higher than ultimate oil recovery.

The reason is that, after dew point conditions are reached, stock-tank oil is continuously being left

behind as immobile condensate trapped in the reservoir. Gas saturation ( ) remains constant

0

10

20

30

40

50

60

70

80

90

100

0 1000 2000 3000 4000 5000

OO

IP R

eco

very

(%

) /

OG

IP R

eco

very

(%

)P

rod

uct

ion

GO

R (

MSC

F/ST

B)

/ G

as S

atu

rati

on

(%

)

Reservoir Pressure (psia)

Gp/G (%) Np/N (%) GOR (MSCF/STB) Sg (%)

Dew Point Pressure

Sg

GOR

Np/N

Gp/G

131

above the dew point because there is no oil phase in the reservoir and water saturation ( ) is

assumed to be constant when no water encroachment is acting on the reservoir system. Gas

saturation ( ) decreases as reservoir pressure decreases around dew point conditions because of

the new presence of reservoir oil phase in the reservoir. Condensate dropout rate is maximum at

conditions near the dew point. When gas condensate is rich enough and pressure is low enough,

gas saturation can slightly increase with reservoir pressure. Gas phase production, oil phase

expansion, and retrograde condensation near to dew point conditions tend to decrease gas

saturation ( ). Expansion of the reservoir gas and re-vaporization of condensate at low pressures

tend to increase gas saturation ( ). Production gas-oil ratio ( ) is constant above the dew

point because composition of produced reservoir gas remains unchanged for such conditions.

Below the dew point, production gas-oil ratio ( ) increases with decreasing reservoir pressure

because of retrograde condensation and the changing nature of the produced reservoir gas.

Production gas-oil ratio ( ) can slightly decreases at low reservoir pressure due to liquid re-

vaporization.

132

5.2.2 Misuse of Gas Condensate Tank Model in Volatile Oil Reservoir

Simulation results from the Simplified Gas Condensate Tank (SGCT) model must be

crosschecked against the typical phase and depletion behavior of gas condensate reservoirs

outlined in the preceding section. If the reservoir fluid is not a gas condensate, prediction results

will be inconsistent with those typical PVT property and depletion behavior. For example, if the

SGCT model is inadvertently used for the analysis of a different type of near-critical fluid, such

as a volatile oil reservoir, calculated gas saturation ( ) profiles will start from zero at saturation

conditions (which would actually represent a bubble point) and monotonically increase with

decreasing reservoir pressure below saturation conditions. Such gas saturation profile is

significantly different from the typical -profile for gas condensates which must starts at one if

no water is present (or at initial gas saturation equal to ) and decreases with decreasing

reservoir pressure at conditions below the dew point. In this case, all simulation results must be

disregarded because the assumptions used for the gas condensate tank model (SGCT) are not

applicable for volatile oil reservoir.

To illustrate the differences between the phase and flow behavior between gas

condensates and volatile oils, reservoir temperature can be manipulated. A volatile oil reservoir

behavior can be obtained using the same fluid characterization and composition presented in

Appendix A, but with a reservoir temperature reduced from 300 F to 190 F. Figure 5-10 plots the

phase envelope of reservoir fluid and the two different reservoir depletion paths for the two

reservoir temperatures under consideration. The depletion path at the reservoir temperature of 300

F represents the path of the gas condensate ( ), while the depletion path at the reservoir

temperature of 190 F represents the path of the volatile oil ( ). Both fluids are near critical

fluids but they are found at the opposite sides of the critical point. Simulation results for the gas

condensate were discussed in Figure 5-1 through Figure 5-9 in the preceding section.

133

Figure 5-10: Phase Envelope and Reservoir Depletion Paths

at Two Different Reservoir Temperatures

The resulting standard PVT properties of the volatile oil, which are calculated using the

PVT model for a gas condensate, are shown in Figure 5-11 and Figure 5-12. The characteristics

of gas formation volume factor ( ), oil formation volume factor ( ), volatilized oil-gas ratio

( ), and solution gas-oil ratio ( ) are similar to the typical characteristics of gas condensates.

However, these results cannot be used and are physically meaningless for an actual volatile oil

because their calculation has been based on assuming that the reservoir oil, which is the main

hydrocarbon phase produced in a volatile oil reservoir, remains immobile in the gas condensate

PVT cell.

0

500

1000

1500

2000

2500

3000

3500

4000

-300 -200 -100 0 100 200 300 400 500

Res

ervo

ir P

ress

ure

(psi

a)

Reservoir Temperature (F)

ReservoirDepletion

Path at 300 F"Gas Condensate"

Critical Point

ReservoirDepletion

Path at 190 F

"Volatile Oil"

134

Figure 5-11: Simulated Gas Formation Volume Factor and Volatilized Oil-Gas Ratio

of Volatile Oil using Gas Condensate PVT Model

Figure 5-12: Simulated Oil Formation Volume Factor and Solution Gas-Oil Ratio

of Volatile Oil using Gas Condensate PVT Model

0

50

100

150

200

250

0.0

2.0

4.0

6.0

8.0

10.0

0 1000 2000 3000 4000 5000

Rv

-V

ola

tiliz

ed

Oil

-Gas

Rat

io (

STB

/MM

SCF)

Bg

-G

as F

orm

atio

n V

olu

me

Fac

tor

(RB

/MSC

F)

Reservoir Pressure (psia)

Bg (RB/MSCF) Rv (STB/MMSCF)

Saturated Pressure

Rv

Bg

0

1000

2000

3000

4000

5000

6000

7000

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

0 1000 2000 3000 4000 5000

Rs

-So

luti

on

Gas

-Oil

Rat

io (

SCF/

STB

)

Bo

-O

il Fo

rmat

ion

Vo

lum

e F

acto

r (R

B/S

TB)

Reservoir Pressure (psia)

Bo (RB/STB) Rs (SCF/STB)

Saturated Pressure

Bo

Rs

135

Production data predictions for the volatile oil, as simulated by the SGCT model using

the PVT properties in Figure 5-11 and Figure 5-12, are shown in Figure 5-13. Characteristics of

cumulative gas recovery ( ), cumulative oil recovery ( ), and production gas-oil ratio

( ) are similar to those simulated for gas condensate but the gas saturation ( ) trend is

significantly different, as discussed above. The typical gas saturation ( ) of condensate should

start at initial gas saturation ( ) at conditions above saturation (dew point) conditions

because there should be no liquid hydrocarbon in that state. However, gas saturation ( ) plotted

in Figure 5-13 approaches zero around the saturation pressure, which indicates that oil saturation

( ) does not approach zero but rather approaches a maximum value ( ). Therefore,

all simulation results must be disregarded.

Figure 5-13: Simulated Production Results of Volatile Oil

using Simplified Gas Condensate Tank Model

0

10

20

30

40

50

60

70

80

90

100

0 1000 2000 3000 4000 5000

OO

IP R

eco

very

(%

) /

OG

IP R

eco

very

(%

)P

rod

uct

ion

GO

R (

MSC

F/ST

B)

/ G

as

Satu

rati

on

(%

)

Reservoir Pressure (psia)

Gp/G (%) Np/N (%) GOR (MSCF/STB) Sg (%)

Saturation Pressure

Sg

GOR

Np/N

Gp/G

136

In short, models developed for gas condensate fluids should not be recklessly used

without proper precautions and crosschecks. A crosschecking process has to be carried out either

before or after having generated the simulation results. Before running the simulation, for

example, the calculated fractions of vapor and liquid phases inside the PVT cell and along the

reservoir depletion path should be analyzed. Molar fraction of vapor phase ( ) for a gas

condensate must approach 1.0 as the saturation pressure is approached because there should be no

liquid at a dew point line, and molar fraction of vapor phase ( ) for a volatile oil must approach

zero as the saturation pressure is approached because there should be no vapor at a bubble point

line. This situation is illustrated in Figure 5-14. After running the simulation, the resulting gas

saturation ( ), for example, should be crosschecked as well.

.

Figure 5-14: Mole Fraction Behavior of Vapor Phase Molar Fraction ( )

for Gas Condensates and Volatile Oils

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0 500 1000 1500 2000 2500 3000 3500 4000

F ng

-Ma

lar

Frac

tio

n o

f V

apo

r P

has

e

Reservoir Pressure (psia)

Fng of Gas Condensate Fng of Volatile Oil

Volatile Oil

Gas Condensate

137

5.3 Field Performance Prediction

Field performance prediction calculations were described in Section 4.5 and have been

developed in a stand-alone VBA application. The model is able to predict pressures, flow rates,

cumulative productions, and expected production time by integrating a zero-dimensional

reservoir model, a flow rate and pressure (IPR/TBR) model, and production constraints together.

The input data set used by the model is presented in Appendix A and the generated results are

discussed in Section 5.3.1. In addition, the appearance of a slightly negative hyperbolic decline

coefficient during the decline period, which is very rare to encounter in conventional decline

curve analysis, is observed from this simulation results. This is discussed in Section 5.3.2. Note

that, in this study, field performance predictions begin at a plateau period because built-up has

been neglected by assuming that the operator has a substantial drilling capacity available to

develop the field.

5.3.1 Field Performance Prediction Results

Predictions for cumulative gas production ( ) and cumulative oil ( ) production vs.

production time ( ) are displayed in Figure 5-15. Cumulative gas production ( ) linearly

increases during plateau period because plateau gas flow rate (

) is being maintained.

Plateau period ends when the 55% target recovery factor is reached. Cumulative gas production at

end of plateau ( ) is equal to 216 BSCF. Cumulative gas production at abandonment

condition ( ) is equal to 317 BSCF, which is equivalent 81% gas recovery factor.

Cumulative oil production ( ) also linearly increases at the beginning because of constant

plateau gas flow rate (

) during plateau period and constant volatilized oil-gas ration ( )

above the dew point. When reservoir pressure goes below the dew point, cumulative oil

138

production ( ) increases at lower rate than cumulative gas ( ) production due to a decreasing

volatilized oil-gas ratio ( ). Cumulative oil production at abandonment condition ( ) is

equal to 37 MMSTB, which is equivalent to a 46% condensate recovery factor. These results

indicate, as expected, that total recovery factor of oil/condensate is significantly less than total

recovery factor of gas in typical depletion operations for gas condensate fluids. This is expected

because a large portion of original oil in place is being left as an immobile condensate phase

inside the reservoir.

Figure 5-15: Cumulative Gas and Oil Production vs. Time

Figure 5-16 plots total gas flow rate ( ) and total oil flow rate ( ) vs. production

time ( ). Total gas flow rate ( ) is maintained at the plateau gas flow rate (

) of 213

MMSCF/D during the plateau period. During the decline period, reservoir pressure is not enough

to maintain the plateau gas flow rate (

), thus total gas flow rate ( ) declines with

0

10

20

30

40

50

60

70

80

0

50

100

150

200

250

300

350

400

0 1 2 3 4 5 6 7

Np

-C

um

ula

tive

Oil

Pro

du

ctio

n (

MM

STB

)

Gp

-C

um

ula

tive

Gas

Pro

du

ctio

n (

BSC

F)

Production Time (Year)

Gp (BSCF) Np (MMSTB)

End of Plateau

Gp

Np

Dew Point

G = 393 BSCFN = 81 MMSTBRvi = 206 STB/MMSCF

139

decreasing reservoir pressure. At abandonment condition, total gas flow rate ( ) is equal to 19

MMSCF/D. Total oil flow rate ( ) is constant at 44 MSTB/D above the dew point. Total oil

flow rate ( ) declines because of decreasing volatilize oil-gas ratio ( ) below the dew point

and because of declining total gas flow rate ( ) during decline period. At low pressures, even

when the volatized oil-gas ratio ( ) slightly increases, total oil flow rate ( ) continues to

decline. This is because increasing volatilized oil-gas ratio ( ) is not enough to compensate for

decreasing total gas flow rate ( ) at that condition. At abandonment conditions, total oil flow

rate ( ) equals 1.3 MSTB/D.

Figure 5-16: Total Gas and Oil Flow Rates vs. Time

Figure 5-17 plots reservoir pressure ( ), bottomhole flowing pressure ( ), and

wellhead pressure ( ) vs. production time ( ). During plateau period, production time ( ) is

directly proportion of cumulative gas recovery ( ) variable because total gas flow rate ( )

0

10

20

30

40

50

0

50

100

150

200

250

0 1 2 3 4 5 6 7

qo

sc-

Tota

l Oil

Flo

w R

ate

(M

STB

/D)

qg

sc-

Tota

l Ga

s Fl

ow

Ra

te (

MM

SCF/

D)

Production Time (Year)

q_gsc (MMSCF/D) q_osc (MSTB/D)

End of Plateau

qgsc

qosc

Dew Point

140

and original gas in place ( ) are constant. Thus, the relationship between reservoir pressure ( )

and production time ( ) during plateau period is similar to the relationship between reservoir

pressure ( ) and cumulative gas recovery ( ) calculated from the gas condensate tank

model. Reservoir pressure ( ) decreases from 4000 psia at initial condition to 3031 psia at the

dew point. After that, reservoir pressure ( ) decreases at a slower rate as a result of the

implementation of a two-phase mode of operation in the gas condensate tank model. This

behavior could be observed in Figure 5-9 as well. In addition, reservoir pressure ( ) during the

decline period decreases at an even slower pace thanks to a declining gas flow rate ( ).

Reservoir pressure ( ) reaches abandonment pressure of 750 psia after 6.23 year of production.

Figure 5-17: Reservoir Pressure, Bottomhole Flowing Pressure,

and Wellhead Pressure vs. Time

0

500

1000

1500

2000

2500

3000

3500

4000

4500

0 1 2 3 4 5 6 7

Pre

ssu

re (p

sia)

Production Time (Year)

Pr (psia) Pwf (psia) Pwh (psia)

End of Plateau

pr

pwf

pwh

Dew Point

141

Figure 5-17 reveals that bottomhole flowing pressure ( ) decreases from 3440 psia at

initial conditions to 668 psia at abandonment conditions. Bottomhole flowing pressure ( )

depends on reservoir pressure ( ) and drawdown pressure ( ), which is the pressure drop

required to produce hydrocarbons out of the reservoir. Wellhead pressure ( ) decreases from

2438 psia at initial condition to minimum allowable wellhead pressure ( ) of 550 psia at end

of plateau. After that, wellhead pressure ( ) is maintained at constant level of 550 psia.

Wellhead pressure ( ) varies with bottomhole flowing pressure ( ) and pressure drop inside

the tubing, which is a pressure drop that must be maintained in the tubing in order to produce

hydrocarbon out of the wellbore.

Drawdown pressure ( ), which is the difference between reservoir pressure ( ) and

bottomhole flowing pressure ( ), is relatively constant above the dew point because most of

the variables in IPR equation (Equation 4-134) are relatively constant. Drawdown pressure ( )

found within the range 545 to 560 psia during this period. Below the dew point, drawdown

pressure ( ) increases with decreasing reservoir pressure ( ) mainly because of decreasing

relative permeability of gas ( ). At reservoir pressures close to the dew point, relative

permeability of gas ( ) decreases significantly due to decreasing gas saturation ( ), while the

product of gas viscosity ( ) times gas formation volume factor ( ) decreases slightly. As a

result, drawdown pressure ( ) has to increase in order to maintain plateau gas flow rate

(

) so that it can compensate for the decreased mobility of the gas phase. As reservoir

pressure continues to decrease, relative permeability of gas ( ) is relatively constant because

gas saturation ( ) is relatively stable (see Figure 5-18), while the product of gas viscosity ( )

times gas formation volume factor ( ) slightly increases. As a result, drawdown pressure ( )

has to continue to increase in order to maintain the plateau gas flow rate (

). During

142

decline period, drawdown pressure ( ) is lower mainly because of declining total gas flow rate

( ).

Figure 5-18: Gas Saturation and Specific Gravity of Reservoir Gas vs. Time

Pressure drop inside the tubing is the difference between bottomhole flowing pressure

( ) and wellhead pressure ( ). Pressure drop inside the tubing is relatively constant above

the dew point because most of the variables in TPR equation (Equation 4-139) are relatively

constant. This pressure drop ranges between the values of 870 to 1000 psi during this above-dew-

point period. During decline period, pressure drop inside the tubing continues to decrease owing

to decreasing total gas flow rate ( ). At abandonment condition, pressure drop inside the

tubing reaches the minimum value of 118 psi.

0.5

0.7

0.9

1.1

1.3

1.5

0.00

0.20

0.40

0.60

0.80

1.00

0 1 2 3 4 5 6 7

Spe

cifi

c G

ravi

ty o

f R

ese

rvo

ir G

as

Gas

Sat

ura

tio

n (

Frac

tio

n)

Production Time (Year)

Gas Saturation SG of Reservoir Gas

End of Plateau

Gas Saturation

Specific Gravity of Reservoir Gas

Dew Point

143

5.3.2 Decline Trend Analysis

Total gas flow rate ( ) and cumulative gas production during decline period (

) is calculated by combining gas condensate tank model and nodal analysis as described

in Section 4.5. These results, for the scenario of interest, are plotted in Figure 5-19. The

exponential decline trend which calculated from exponential decline equation is also plotted into

the same figure. Total gas flow rate ( ) from exponential decline trend is calculated from

Equation 5-2 and the decline rate ( ) of exponential decline trend is calculate from Equation

4-144.

Equation 5-2

Figure 5-19: Total Gas Flow Rate ( ) vs.

Cumulative Gas Production during Decline Period

0

50

100

150

200

250

0 20 40 60 80 100 120

qgs

c-

Tota

l Gas

Flo

w R

ate

(M

MSC

F/D

)

(Gp - Gpplateau) - Cumulative Gas Production during Decline Period (BSCF)

Field Performance Preidction Data Exponentail Decline Trend (b = 0.0)

Exponential Decline TrendDi = 1.92 * 10-3 {Day-1}

Field Performance Prediction Data

144

From Figure 5-19, calculated results from field performance prediction and exponential

decline equation agree with each other, which is not surprising. The assumptions used in this

simulator, including pseudo steady state flow condition, no water production, no water injection,

no water influx, and constant wellhead pressure during decline period, are favorable assumptions

for an exponential decline. Therefore, the exponential decline is a fairly good assumption for

calculating production time ( ) based on total gas flow rate ( ) and cumulative gas production

( ).

However, detailed analysis of the decline trend shows decline rate ( ) actually varies -

although very slightly. If exponential decline is assumed between each decline interval, decline

rate ( ) of each interval can be calculated from Equation 5-3. The calculation results which are

plotted in Figure 5-20 show that decline rate ( ) increases with increasing cumulative gas

production ( ) for most of the time. In other words, they suggest that decline rates are slightly

increasing in time, which in turn implies having a hyperbolic decline exponent ( ) of negative

value. In conventional decline curve analysis, decline rates ( ) are always expected to dampen in

time and thus hyperbolic decline coefficients ( ) are always expected to be positive (

). A negative value for the decline exponent ( ) is extremely rare scenario. It became

apparent during this study that this decline behavior was coupled with the appearance/

disappearance of the condensate phase and related property changes during decline, as suggested

by Figure 5-20. In this figure, decline rates slightly increase with reservoir production but they

reach a maximum after which they start to decrease. This event seems to closely follow the

behavior of GOR presented in Figure 5-9. It is suggested that condensate effects and the inherent

evolution of fluid properties in time impose the increase in the decline rate at earlier times during

the decline period but this trend is reversed around the moment GOR reaches its maximum. It is

important to note that these decline rate changes are not very large or significant; therefore the

exponential decline assumption still remains largely valid for engineering evaluation purposes. It

145

should also stressed that conventional decline curve analysis is based on the assumption of

production at a constant bottomhole pressure ( ) which is not strictly valid for the scenario

under consideration as displayed in Figure 5-17 during decline.

Equation 5-3

Figure 5-20: Decline Rate ( ) vs. Cumulative Gas Production

during Decline Period ( )

1.6E-03

1.7E-03

1.8E-03

1.9E-03

2.0E-03

2.1E-03

2.2E-03

0 20 40 60 80 100 120

D -

De

clin

e R

ate

(1/D

ay)

(Gp - Gpplateau) - Cumulative Gas Production during Decline Period (BSCF)

Decline Rate (1/Day)

Decline Rate of 1.92 * 10-3 {Day-1}

146

5.4 Economic Analysis and Optimization

An economic evaluation model has been developed to calculate net present value (NPV)

and rate of return (ROR) based on the simplified economic model described in Section 4.6.1.

Results from field performance prediction from the previous section and economic parameters

from Appendix A are inputted into this economic model, and the generated results are shown and

discussed in Section 5.4.1. For field optimization studies, results of the sensitivity analysis of

NPVs for different target recovery at end of plateau and total number of wells are displayed and

elaborated upon in Section 5.4.2.

5.4.1 Field Economic Analysis

Figure 5-21 presents the predictions for annual expenditure, annual revenue, and

cumulative discounted net cash flow vs. production time for the reservoir exploitation scenario

under consideration. Annual expenditure is equal to 1480 Million $ at the beginning. It is very

large because all CAPEX, including drilling and completion cost, platform cost, pipeline cost,

and production facilities cost, is spent at that time. Annual expenditure drops drastically to 116

Million $ in the 1st year of production because it consists of OPEX and taxes only and they are

relatively small when compared to initial CAPEX. Annual expenditure continuously decreases

because ad valorem and severance taxes decrease resulting from decreasing in annual revenue.

Annual expenditure at the last year of production is equal to 8.21 Million $. Annual revenue is

equal to 1373 Million $ in the 1st year of production. Then, it continuously declines due to

decreasing gas and oil flow rates. Annual revenue at the last year of production is equal to15.70

Million $.

147

Figure 5-21: Annual Expenditure, Annual Total Revenue,

and Cumulative Discounted Net Cash Flow vs. Production Time

Figure 5-21 shows that the cumulative discounted net cash flow is equal to (-1480)

Million $ at the beginning of the project. Cumulative discounted net cash flow increases every

year because annual revenue is higher than annual expenditures throughout the production period.

However, the rate of net cash increase decreases at late time because net cash flow is lower and

time discount factor is higher. The cumulative discounted net cash flow at the last year of

production, which is equivalent to the project’s NPV, is equal to 1189 Million $. Figure 5-21 also

shows that cumulative discounted net cash flow turns from negative value to positive value

between the first year and the second year of production. Thus, the payback period, which is the

time period required for cumulative discounted net cash flow to be equal to zero, is between one

to two years.

(2000)

(1500)

(1000)

(500)

0

500

1000

1500

2000

0 1 2 3 4 5 6 7

Mo

ne

tary

Val

ue

(M

illio

n $

)

Production Period (Year)

Annual Expenditure Annual Revenue Cumulative Discounted Net Cash Flow

Annual Expenditure

Annual Revenue

Cumulative Discounted Net Cash Flow

148

Figure 5-22: Net Present Value vs. Interest Rate

Figure 5-22 shows project’s NPVs at different interest or discount rates ranging from 5%

to 60%. The NPV decreases with increasing interest or discount rate because future net cash flow

is more heavily discounted and penalized. The discount interest rate has less impact on project’s

expenditures than on its revenue because most of the investment is spent at the beginning of the

project while most of the revenue is actually generated later in time. The NPV is equal to 1590

Million $ at interest rate of 5% and monotonically decreases to 245 Million $ at interest rate of

40%. Rate of return (ROR), which is the interest rate that results in zero NPV, is equal to 53%.

(500)

0

500

1000

1500

2000

0.00 0.10 0.20 0.30 0.40 0.50 0.60

NP

V -

Net

Pre

sen

t Val

ue

(Mill

ion

$)

Interest Rate (Fraction)

ROR = 53%NPV @ 12% Interest Rate = 1189 Million $

149

5.4.2 Field Optimization

Figure 5-23 is a composite figure that shows field optimization results in both tabular and

graphical forms. Target recovery factors at end of plateau are varied between 30% and 75% while

total number of wells required for development is varied between 3 and 30 wells. It is readily

realized from this figure that the optimized NPV can be placed at 1326 Million $ for the

combination of 30% target recovery factor and the use of 15 wells for field development.

Figure 5-23: Field Optimization Results

3

12

21

30

0

200

400

600

800

1,000

1,200

1,400

0.300.40

0.500.60

0.70

TotalNumberof Wells

Net

Pre

sen

t V

alu

e (M

illio

n $

)

Target Recovery Factor at End of Plateau

NPV Target Recovery Factor at End of Plateau

(Million $) 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65 0.70 0.75

3 639 601 550 482 394 281 136 -50 -283 -579

6 1,077 1,052 1,016 966 895 801 667 480 210 -194

9 1,244 1,226 1,199 1,161 1,106 1,028 917 753 499 76

12 1,310 1,296 1,275 1,244 1,199 1,134 1,040 899 670 262

15 1,326 1,315 1,298 1,273 1,233 1,179 1,098 975 771 388

18 1,313 1,304 1,290 1,269 1,238 1,189 1,120 1,010 827 469

21 1,284 1,277 1,264 1,247 1,217 1,177 1,117 1,018 852 521

24 1,241 1,236 1,225 1,209 1,186 1,151 1,096 1,010 856 550

27 1,195 1,190 1,182 1,167 1,147 1,113 1,065 985 847 561

30 1,139 1,136 1,129 1,115 1,097 1,070 1,024 954 827 559

Tota

l Nu

mb

er o

f W

ell

150

If a target plateau recovery factor is fixed due to contractual obligations or market

saturation demands, optimum number of wells could be determined. When too few wells are

drilled to develop a hydrocarbon deposit, resulting field flow rates will be smaller and the

required production period needed to reach abandonment will be prolonged. Thus, annual revenue

from a distant future will be dramatically discounted, which will result in lower NPVs. In

contrast, if total number of wells used to develop the reservoir is too high, the additional revenue

that would be obtained from production acceleration will not be able to compensate or offset the

significantly increased drilling and completion costs. As a result, NPV will not be maximized

under either scenario.

When the total number of wells is fixed, an optimum target recovery factor could not be

found. Under the current model, NPV will always increases with decreasing target recovery

factor. This is because CAPEX does not change with total flow rates in this economic model. If

target recovery factor is lowered, initial flow rates will be higher and annual revenue will be

accelerated. However, production facilities costs, required to handle such an increased volume of

fluids, remain constant in this study. As a result, NPV will always be better for lower target

plateau recovery factor because the production is accelerated and the revenue can be received

earlier in the life of the field. In actual field applications, there is always a maximum fluid volume

that can be reasonably handled at the surface and accepted by the market. This justifies the need

for a plateau period. In the limiting case when target plateau recovery becomes zero, which

effectively eliminates the plateau period, initial flow rates for the decline period will be extremely

large. This is good news for the simplified economic model but bad news in real applications

because those fluid volumes might not be able to be marketed effectively and the required surface

facilities would become extremely expensive. In addition, facilities designed to handle such large

volumes just during the first year alone would become awfully overdesigned for the rest of the

reservoir life - a situation that is far from optimal. In order to be able to actually determine a more

151

realistic and optimum target recovery factor, the CAPEX model has to be adjusted. Surface

production facilities cost and/or flowline costs have to be made functions of maximum expected

flow rates by introducing either a continuous- or step- cost function in the CAPEX model.

152

5.5 Application for Other Production Situations

This section discusses recommendations on how to extend the capabilities of field

performance simulator for gas condensate fluids in order to tackle other possible production

scenarios. Two other scenarios, such as dry gas / wet gas production and gas condensates with

producible reservoir oil, are elaborated upon. The fundamental differences and the required

modifications to current gas condensate scenario build in the present model, as well as the

expected results from such modified models, are the main areas of discussion in this section.

153

5.5.1 Application for Dry Gas / Wet Gas

Even though the proposed model has been specifically tailored to the analysis of gas

condensate reservoir fluids, it could be actually be applied to any other natural gas reservoir, such

as dry gases and wet gases, with few modifications. In dry gas and wet gas reservoirs, the

hydrocarbon fluid is always found in a 100% vapor phase state throughout their isothermal

reservoir depletion path. One of the main differences between dry gas and wet gas is that, along

the surface depletion path, dry gas will also stay in a 100% vapor phase condition, while wet gas

will experience two-phase condition or condensate dropout as it flows through the surface

production system.

The procedure used to calculate standard PVT properties of gas condensate at the dew

point could be applied for the dry gas and wet gas. Characteristic of gas formation volume factor

( ) is expected to be the same for dry gases, wet gases, and gas condensates. Volatilized oil-gas

ratio ( ) for a dry gas will be zero, while volatilized oil-gas ratio ( ) of the wet gas will be

constant. Oil formation volume factor ( ) and solution gas-oil ratio ( ) will not be defined or

calculated in dry gas and wet gas because there is no presence of reservoir oil phase along the

reservoir depletion path. In addition, specific gravity of reservoir gas will be constant for both dry

gas and wet gas.

For the reservoir zero-dimensional model, the gas condensate tank model above the dew

point could be applied directly to dry gases and wet gases. However, because there is only

surface-gas in the reservoir gas phase, cumulative oil recovery ( ) and original oil in place

( ) for a dry gas will be zero. Based on the same assumptions as gas condensate tank model, gas

saturation ( ) in dry gas and wet gas will be constant at initial gas saturation ( ).

In IPR calculations, relative permeability of gas ( ) will be constant because gas

saturation is constant. Gas flow rate ( ) will depend on drawdown pressure ( ) and the

154

multiplication of gas viscosity ( ) and gas formation volume factor ( ). Oil flow rate ( ) for

dry gas will be zero because volatilized oil-gas ratio ( ) is zero. Oil flow rate ( ) for wet gas

will be equal to gas flow rate ( ) times volatilized oil-gas ratio ( ) which is constant. In TPR

calculations, pressure drop inside the tubing of dry gas and wet gas will be relatively stable

comparing to gas condensate because specific gravity of reservoir gas are constant. For the case

of the dry-gas, there is no liquid drop out in the tubing.

Field performance prediction, economic analysis, and field optimization for dry gases and

wet gases are carried out using the same procedure as for a gas condensate above the dew point.

Cumulative gas recovery ( ) at abandonment condition of dry gas is expected to be higher

than wet gases and gas condensates because lighter hydrocarbons exhibit larger expansivity

coefficients or isothermal compressibility values and there is no obstruction to fluid flow due to

the presence of a liquid hydrocarbon phase. Cumulative oil recovery ( ) at abandonment

condition for wet gases is expected to be much higher than gas condensate because the reservoir

system does not leave any immobile condensate inside the reservoir. In term of NPV, dry and wet

gases are expected to generate less NPV than gas condensates because they produce much less oil

which can be sold for a much higher commodity price. In terms of optimization of target recovery

factor at end of plateau and total number of wells, dry and wet gases are expected to exhibit

similar characteristics as those of gas condensates.

155

5.5.2 Application for Gas Condensate with Producible (Mobile) Reservoir Oil

One of the significant assumptions of the gas condensate tank model is that reservoir oil

phase is immobile. However, this assumption is not always valid, especially around the wellbore

where oil saturation might build up to a high enough value so that the relative permeability to oil

might not be equal to zero anymore. This situation would increase complexity of gas condensate

system.

Standard PVT properties simulated from procedures described in either section 4.2.2 or

4.2.3 will carry certain degree of error toward the final results because those algorithms assumed

that the reservoir oil remained immobile. However, there is no practical approach for modifying

those procedures in order to fully represent the producible reservoir oil scenario. In order to

simulate the producible reservoir oil situation within a standard PVT properties calculation

algorithm, the ratio between excess gas and excess oil which should be removed from the PVT

cell during the constant volume expansion has to be given. This ratio depends on the expected

relative mobility ratio between gas and oil phases which are the functions of their relative

permeabilities and fluid properties. Relative permeability depends on saturation fraction and

saturation fraction could be obtained from saturation equation; thus, relative permeability is also

the function of fluid properties. Moreover, the shape of the relative permeability curves is also

rock-dependent and not solely fluid property dependent. In short, there is no simple method to

fully and reliably represent producible reservoir oil situation.

In the reservoir zero-dimensional model, the gas condensate tank model has to be

modified by substituting Equation 4-115 with Equation 5-4 (below), substituting Equation 4-118

with Equation 5-5 (below), and calculating instantaneous production GOR at pressure level j ( )

using Equation 5-6 (below). Relative permeability in Equation 5-6 might be evaluated from

156

saturation value calculated at the preceding pressure level j-1. The definition of each parameter in

these equations can be found in the gas condensate tank model section.

Equation 5-4

Equation 5-5

Equation 5-6

For IPR calculations, gas flow rate ( ) and oil flow rate ( ) should be calculated by

implementing Equation 4-132 and Equation 4-133 instead. For TPR calculations, the single phase

gas flow equation is still applied; but pipe efficiency factor ( ) is expected to be lower

because of additional liquid phase flow into wellbore.

Procedures used to perform field performance prediction, economic analysis, and field

optimization remain unchanged. Cumulative gas recovery ( ) at abandonment condition for

producible reservoir oil scenario is expected to be close to the value obtained from immobile

reservoir oil scenario. In contrast, cumulative oil recovery ( ) at abandonment condition for

producible reservoir oil scenario is expected to be higher than the value obtained from immobile

reservoir scenario because less oil is left immobile inside the reservoir. In term of NPV

157

estimations, larger NPVs are expected from producible reservoir oil scenario because reservoir oil

could yield more stock-tank oil, which is more expensive product, than reservoir gas. For field

optimization, the similar characteristic is expected whether reservoir oil is producible or not.

Chapter 6

Summary and Conclusions

A model able to perform field performance analysis and optimization of exploitation

strategies for a gas condensate reservoir has been successfully developed. The model has been

constructed using Microsoft Excel with built-in Visual Basic for Applications (VBA) program.

The model includes a fluid property calculation subroutine which estimates standard PVT

properties based on a binary pseudo-component model using fluid compositional information as

input. The subroutine demonstrates to produce reliable standard PVT properties typical of gas

condensate fluid phase behavior. Limitations of the pseudo-component model, such as the

generation of negative solution gas-oil ratio ( ) at low reservoir pressure, are clearly shown and

explained using the simulated results produced by the fluid property calculation subroutine.

A zero-dimensional reservoir model based on the generalized material balance equation

for gas condensates has also been developed. The results from this gas condensate tank model are

able to mimic the typical reservoir performance data found in gas condensate fields. Possible

sources of error and misinterpretation from using a gas condensate tank model in the analysis of

other near critical fluids, such as volatile oils, are discussed and recommended crosschecking

procedure that should be implemented before and after running the simulation model is

addressed.

A field performance prediction that couples zero-dimensional reservoir models with

nodal analysis concepts has been successfully developed for the screening of field development

strategies. The field performance prediction tool has also been coupled with an economic model

which enables the prediction of optimum field development strategies. The reservoir model

demonstrates to provide results which are consistent with reservoir depletion behavior for gas

159

condensates. A discussion of the observed decline trend analysis has been included to shed some

light on the possibility of obtaining slightly negative hyperbolic decline coefficients in gas

condensate fields. The economic evaluation subroutine was implemented based on simplified

economic model. The subroutine is used to optimize target development variables, such as target

recovery factor at end of plateau and total number of wells required to optimally develop the

field. This proposed model is shown to be able to perform economic analysis and field

optimization effectively. Economic and optimization results are analyzed in detail. Limitations of

the economic model assumptions are addressed and discussed.

The proposed model is suitable for real field applications when either input data or

working time is constrained. For example, this model is appropriate to be used to simulate field

performance data for feasibility study of gas condensate reservoirs because, during that phase of

field development, reservoir data is usually limited and all available data are highly uncertain;

thus constructing highly sophisticated model is impractical. In addition, this model can effectively

simulate field performance data for numerous production scenarios, which is a very important

factor to cope with the high uncertainty found in that period. Another proper application is to use

this proposed model to perform project evaluations for new asset acquisitions because, during the

acquisition process, the evaluation of each project has to be completed within a short period of

time; thus utilizing the less complicated model, which takes less time to construct and execute, is

more feasible, even if there are plenty of reservoir and field development data.

Additional recommendations for avenues for future work are also provided for the

improvement of the reliability and the capabilities of the proposed field performance model. The

first recommendation is to validate the accuracy of the model’s outputs with the outputs

calculating from full scale full dimensional numerical simulator using both hypothetical data and

actual field data. With the hypothetical data validation, the error resulting from a single-

homogeneous tank assumption (neglecting all gradients) can be analyzed. With the actual field

160

data validation, the impact of heterogeneity on naturally occurring reservoirs can be determined.

The second recommendation is to perform sensitivity analysis to investigate how the variation in

economic outputs can be attributed to the variation in reservoir and field development input

variables. With this understanding, the limited resources (time and man-hour) can be properly

spent to evaluate the expected values and uncertainties of the variables with highly economic

impact. In most cases, product prices, development costs, and reserves are expected to be the

variables with the highest impact on the project’s economic. The other recommendations include

the handling of production from dry gas and wet gas reservoirs and the modeling of gas

condensates with producible (mobile) reservoir oil. Further avenues include allowing the model

to handle multiple reservoirs and multiple types of producing wells, which would make the model

applicable to be used in complex reservoir structures and marginal fields development.

Bibliography

Ayala, L.F. 2009a. Lecture Notes, PNG530 Natural Gas Engineering, Department of Energy and

Mineral Engineering, Pennsylvania State University.

Ayala, L.F. 2009b. Phase Behavior Model Instruction, PNG520 Phase Behavior of Petroleum

Fluids, Department of Energy and Mineral Engineering, Pennsylvania State University.

Baker, L.E., Pierce, A.C., and Luks, K.D. 1982. Gibbs Energy Analysis of Phase Equilibria.

SPEJ. 22 (5): 731-742

Coats, K.H. 1985. Simulation of Gas Condensate Reservoir Performance. Paper SPE 10512,

Journal of Petroleum Technology. October: 1870-1886.

Colebrook, C.F. 1939. Turbulent Flow in Pipes, with Particular Reference to the Transition

Region Between the Smooth and Rough Pipe Laws. Journal of the Institution of Civil Engineers.

February: 133-156

Danesh, A. 1998. PVT and Phase Behavior of Petroleum Reservoir Fluids: Elsevier

Deitz, D.N. 1965. Determination of Average Reservoir Pressure from Build-Up Surveys. Journal

of Petroleum Technology. 17 (8): 955-959

Ertekin, T., Abou-Kassem, J.H., and King, G.R. 2001. Basic Applied Reservoir Simulation, SPE

Textbook Series Vol.7: Society of Petroleum Engineer

Jhaveri, B.S. and Youngren, G.K. 1988. Three-Parameter Modification of the Peng-Robinson

Equation of State to Improve Volumetric Predictions. SPERE. 3 (3): 1033-1040

Jossi, J.A., Stiel, L.I., and Thodos, G. 1962. The Viscosity of Pure Substances in the Dense

Gaseous and Liquid Phases. AIChE Journal. 8 (1): 59-63

Lee, A.L., Gonzalez M.H., and Eakin, B.E. 1966. The Viscosity of Natural Gases. Journal of

Petroleum Technology. 18 (8): 997-1000

Lee, J., Rollins, J.B., and Spivey J.P. 2003. Pressure Transient Testing, SPE Textbook Series

Vol.9: Society of Petroleum Engineer

Lohrenz, J., Bray, B.G., and Clark, C.R. 1964. Calculating Viscosities of Reservoir Fluids from

Their Compositions. Journal of Petroleum Technology. 16 (10): 1171-1176

Martin, J.J. 1979. Cubic Equations of State-Which?. Industrial and Engineering Chemical

Fundamentals. 18 (2): 81-97

McCain, W.D., Jr. 1990. The Properties of Petroleum Fluid, 2nd

Ed: PennWell

Mehra, R. K., Heidemann, R. A. and Aziz, K. 1983. An Accelerated Successive Substitution

Algorithm. The Canadian Journal of Chemical Engineering. 61 (4): 590-593

162

Mian, M.A. 2002. Project Economics and Decision Analysis, Volume I: Deterministic Models:

PennWell

Michelsen, M.L. 1982a. The Isothermal Flash Problem. Part I. Stability. Fluid Phase Equilibria.

9 (1): 1-19

Michelsen, M.L. 1982b. The Isothermal Flash Problem. Part II. Phase-Split Calculation. Fluid

Phase Equilibria. 9 (1): 21-40

Peneloux, A., Rauzy E., and Freze R. 1982. A Consistent Correction for Redlich-Kwong-Soave

Volumes. Fluid Phase Equilibria. 8 (1): 7-23

Peng, D.Y. and Robinson, D.B. 1976 A New Two-Constant Equation of State. Industrial and

Engineering Chemical Fundamentals. 15 (1): 59-64

Rachford, H.H. and Rice, J.D. 1952. Procedure for Use of Electronic Digital Computers in

Calculating Flash Vaporization Hydrocarbon Equilibrium. Papar SPE 952327-G, Petroleum

Transaction, AIME 195: 327-328

Redlich, O. and Kwong, J.N.S. 1949. On the Thermodynamics of Solutions. V. An Equation of

State. Fugacities of Gaseous Solutions. Chemical Reviews. 44 (1): 233-244

Soave, G. 1972. Equilibrium Constants from a Modified Redlich-Kwong Equation of State.

Chemical Engineering Science 27 (6): 1197-1203

Thararoop, P. 2007. Computer Project No.1: Analysis of Gas Condensate Reservoir Exploitation,

PNG530 Natural Gas Engineering, Department of Energy and Mineral Engineering, Pennsylvania

State University.

Walkenbach, J. 2007. Excel 2007 Power Programming with VBA. Wiley Publishing, Inc

Walsh, M.P. 1995. A Generalized Approach to Reservoir Material Balance Calculations. Journal

of Canadian Petroleum Technology. 34 (1)

Walsh, M.P. and Lake, L.W. 2003. A Generalized Approach to Primary Hydrocarbon Recovery:

Elsevier

Walsh, M.P. and Towler, B.F. 1995. Method Computes PVT Properties for Gas Condensates. Oil

and Gas Journal. July: 89-83

Wilson, G. 1968. A Modified Redlich-Kwong EOS, Application to General Physical Data

Calculations. Paper No 15C presented at the AIChE 65th National Meeting

Whitson, C.H. and Brule, M.R. 2000. Phase Behavior, SPE Monograph Series Volume 20.

Society of Petroleum Engineer

Appendix A

Input Data Summary

Table A-1: Pressures and Temperatures for

Standard PVT Properties Calculation Subroutine

Pressure Temperature

(psia) (F)

Reservoir Condition 4000 300

1st Stage Separator 500 90

2nd Stage Separator 150 65

Stock Tank Condition 100 60

Table A-2: Physical Properties of Pure Components

Component Mole

Fraction

Critical

Pressure

(psia)

Critical

Temperature

(R)

Acentric

Factor

Molecular

Weight

(lbm/lbmol)

Critical

Volume

(ft3/lbm)

N2 0.0223 493.10 227.49 0.037 28.013 0.0510

C1 0.6568 666.40 343.33 0.010 16.043 0.0988

C2 0.1170 706.50 549.92 0.098 30.070 0.0783

C3 0.0587 616.00 666.06 0.152 44.097 0.0727

i-C4 0.0127 527.90 734.46 0.185 58.123 0.0714

n-C4 0.0168 550.60 765.62 0.200 58.123 0.0703

i-C5 0.0071 490.40 829.10 0.228 72.150 0.0679

n-C5 0.0071 488.60 845.80 0.251 72.150 0.0675

n-C6 0.0138 436.90 913.60 0.299 86.177 0.0688

n-C10 0.0832 305.20 1112.00 0.490 142.285 0.0679

CO2 0.0045 1071.00 547.91 0.267 44.010 0.0344

164

Table A-3: Binary Interaction Coefficients of Pure Components

δij's N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 n-C6 n-C10 CO2

N2 0.0000 0.0180 0.0390 0.0460 0.0470 0.0470 0.0480 0.0480 0.0000 0.0000 0.0000

C1 0.0180 0.0000 0.0050 0.0100 0.0145 0.0145 0.0182 0.0182 0.0000 0.0000 0.0000

C2 0.0390 0.0050 0.0000 0.0017 0.0032 0.0032 0.0048 0.0048 0.0000 0.0000 0.0000

C3 0.0460 0.0100 0.0017 0.0000 0.0012 0.0012 0.0024 0.0024 0.0000 0.0000 0.0000

i-C4 0.0470 0.0145 0.0032 0.0012 0.0000 0.0000 0.0008 0.0008 0.0000 0.0000 0.0000

n-C4 0.0470 0.0145 0.0032 0.0012 0.0000 0.0000 0.0008 0.0008 0.0000 0.0000 0.0000

i-C5 0.0480 0.0182 0.0048 0.0024 0.0008 0.0008 0.0000 0.0000 0.0000 0.0000 0.0000

n-C5 0.0480 0.0182 0.0048 0.0024 0.0008 0.0008 0.0000 0.0000 0.0000 0.0000 0.0000

n-C6 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

n-C10 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

CO2 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

Table A-4: Volume Translation Coefficient of Pure Components

Component Si for PR

EOS

Si for SRK

EOS

N2 -0.19270 -0.00790

C1 -0.15950 0.02340

C2 -0.11340 0.06050

C3 -0.08630 0.08250

i-C4 -0.08440 0.08300

n-C4 -0.06750 0.09750

i-C5 -0.06080 0.10220

n-C5 -0.03900 0.12090

n-C6 -0.00800 0.14670

n-C10 0.06550 0.20800

CO2 -0.08170 0.08330

165

Table A-5: Reservoir Input Data

Reservoir

Number

Initial

Reservoir

Pressure

(psia)

Abandonment

Reservoir

Pressure

(psia)

Dew Point

Reservoir

Pressure

(psia)

Reservoir

Temperature

(Deg F)

Temperature

Gradient

(Deg F / ft)

1 4000 750 3031 300 0.0205

Temperature

Surface

(Deg F)

Reservoir

Depth

(ft)

Reservoir

Drainage

Area

(Acres)

Reservoir

Thickness

(ft)

Reservoir

Pososity

(Frac)

Connate

Water

Saturation

(Frac)

60 6560 5000 55 0.200 0.210

Original Gas

Equivalent

In Place

(BSCF)

Absolute

Permeability

(md)

Wellbore

Radius

(ft)

Shape Factor Mechanical

Skin

Non-Darcy

Flow Coeff

(D/MSCF)

393.15 80 0.375 30.88 10 0.0011

Table A-6: Relative Permeability Input Data

Reservoir

Number

Connate

Water

Saturation

Connate

Gas

Saturation

Maximum

Gas

Saturation

Krg at

Maximum

Gas

Saturation

1 0.20 0.20 0.80 0.80

166

Table A-7: Standard PVT Properties

Pressure Bo Bg Rs Rv SG Gas

(psia) (RB/STB) (RB/MSCF) (SCF/STB) (STB/MMSCF)

4000 5.052 1.040 4856.8 205.9 1.147

3950 5.089 1.048 4856.8 205.9 1.147

3900 5.127 1.056 4856.8 205.9 1.147

3850 5.166 1.064 4856.8 205.9 1.147

3800 5.206 1.072 4856.8 205.9 1.147

3750 5.248 1.081 4856.8 205.9 1.147

3700 5.291 1.089 4856.8 205.9 1.147

3650 5.336 1.099 4856.8 205.9 1.147

3600 5.382 1.108 4856.8 205.9 1.147

3550 5.429 1.118 4856.8 205.9 1.147

3500 5.479 1.128 4856.8 205.9 1.147

3450 5.530 1.139 4856.8 205.9 1.147

3400 5.583 1.149 4856.8 205.9 1.147

3350 5.638 1.161 4856.8 205.9 1.147

3300 5.695 1.173 4856.8 205.9 1.147

3250 5.754 1.185 4856.8 205.9 1.147

3200 5.815 1.197 4856.8 205.9 1.147

3150 5.879 1.211 4856.8 205.9 1.147

3100 5.946 1.224 4856.8 205.9 1.147

3050 6.015 1.239 4856.8 205.9 1.147

3031 6.042 1.244 4856.8 205.9 1.147

3000 2.858 1.251 2067.5 195.4 1.128

2950 2.706 1.265 1903.7 181.4 1.102

2900 2.582 1.280 1768.5 170.0 1.081

2850 2.479 1.297 1652.8 160.3 1.062

2800 2.389 1.315 1551.4 151.9 1.046

2750 2.311 1.335 1461.0 144.5 1.032

2700 2.241 1.356 1379.3 137.9 1.019

2650 2.178 1.378 1304.9 131.9 1.007

2600 2.120 1.401 1236.4 126.4 0.996

2550 2.067 1.426 1173.1 121.4 0.986

2500 2.019 1.452 1114.1 116.8 0.977

2450 1.974 1.480 1059.0 112.6 0.969

2400 1.932 1.509 1007.3 108.6 0.961

2350 1.892 1.539 958.5 105.0 0.954

2300 1.855 1.571 912.5 101.5 0.947

167

Table A-7: Standard PVT Properties (Cont.)

Pressure Bo Bg Rs Rv SG Gas

(psia) (RB/STB) (RB/MSCF) (SCF/STB) (STB/MMSCF)

2250 1.821 1.605 868.9 98.3 0.941

2200 1.788 1.641 827.5 95.3 0.935

2150 1.757 1.678 788.1 92.5 0.929

2100 1.728 1.718 750.5 89.9 0.924

2050 1.700 1.759 714.6 87.4 0.919

2000 1.673 1.803 680.3 85.1 0.914

1950 1.648 1.850 647.4 82.9 0.910

1900 1.623 1.899 615.9 80.8 0.906

1850 1.600 1.952 585.6 78.9 0.902

1800 1.578 2.007 556.5 77.1 0.899

1750 1.556 2.066 528.4 75.4 0.896

1700 1.536 2.128 501.4 73.8 0.893

1650 1.516 2.195 475.4 72.3 0.890

1600 1.497 2.266 450.3 71.0 0.887

1550 1.479 2.342 426.1 69.7 0.885

1500 1.461 2.424 402.7 68.6 0.883

1450 1.445 2.511 380.1 67.6 0.881

1400 1.428 2.606 358.2 66.6 0.880

1350 1.412 2.707 337.1 65.8 0.879

1300 1.397 2.817 316.6 65.1 0.878

1250 1.382 2.936 296.8 64.5 0.877

1200 1.368 3.065 277.6 64.1 0.876

1150 1.354 3.207 259.0 63.7 0.876

1100 1.341 3.361 241.0 63.5 0.876

1050 1.328 3.532 223.6 63.5 0.877

1000 1.315 3.720 206.8 63.6 0.877

950 1.303 3.928 190.4 63.9 0.879

900 1.291 4.161 174.6 64.3 0.880

850 1.280 4.423 159.3 65.0 0.882

800 1.269 4.719 144.3 66.0 0.885

750 1.257 5.056 129.2 67.3 0.888

700 1.246 5.444 114.5 68.9 0.892

650 1.236 5.894 100.5 71.0 0.897

600 1.226 6.423 86.9 73.5 0.903

550 1.216 7.053 73.9 76.8 0.910

500 1.206 7.816 61.5 80.8 0.919

168

Table A-8: Tubing Input Data

Min Allow Wellhead Pressure (psia) 550

Tubing Roughness (inch) 0.0018

Tubing Efficiency (Frac) 0.70

Depth

(MD - ft)

Depth

(TVD - ft)

Tubing ID

(inch)

Temperature

(F)

0 0 60.0

650 650 4.5 83.8

1300 1300 4.5 107.6

1950 1950 4.5 131.3

2600 2600 4.5 155.1

3250 3250 4.5 178.9

3900 3900 4.5 202.7

4550 4550 4.5 226.5

5200 5200 4.5 250.2

5850 5850 4.5 274.0

6500 6500 4.5 297.8

Table A-9: Economic Input Data

General

Net Hydrocarbon Fraction 87.50%

Price

First Year Gas Price 4.000 $/MSCF

Gas Price Escalation 1.00%

First Year Oil Price 80.000 $/STB

Oil Price Escalation 2.00%

Discount Rate

Discount Rate 12.00%

169

Table A-9: Economic Input Data (Cont.)

Capital Expenses

Capex Total 1,480.0 Million $

- Drilling (per well) 25.0 Million $

- Flowlines and Trunklines 100.0 Million $

- Production Facilities 500.0 Million $

- Platform (for Offshore) 400.0 Million $

- Others 30.0 Million $

Operating Expenses

First Year Opex 0.5 Million $/Mth

Opex Escalation 2.50%

Tax

Gas Severance Tax 8.00%

Oil Severance Tax 8.00%

Ad Valorem Tax 0.00%

Table A-10: Field Performance Prediction Input

Target Recovery

at End of Plateau

Total Number

of Wells

0.55 18

Table A-11: Field Performance Optimization Input

Target Recovery

at End of Plateau

Total Number

of Wells

0.30 3

0.35 6

0.40 9

0.45 12

0.50 15

0.55 18

0.60 21

0.65 24

0.70 27

0.75 30

170

Appendix B

User Guide

1. Open ―Field Development Plan.xlsm

2. Simulate standard PVT properties

a. Simulate standard PVT propertied from Walsh-Tolwer algorithm

i. Select worksheet ―PVT from CVD‖

ii. Input data in ―Reservoir Input‖ section

iii. Input data in ―Pressure Volume Relation‖ section

iv. Input data in ―Constant Volume Depletion‖ section

v. Input data in ―Calculated Cumulative Recovery‖ section

vi. Input data in ―Z-Factor of Produced Wellstreams‖ section

vii. Input data in ―Composition of Produced Wellstreams‖ section

viii. Adjust formulas in ―Pre-Calculation‖ section

ix. Adjust formulas in ―Walsh-Towler Algorithm‖ section

x. Adjust formulas in ―Standard PVT Properties at Dew Point and Below‖

section

xi. Adjust formulas in ―Standard PVT Properties Above Dew Point‖ section

xii. Copy all calculated standard PVT properties to ―FDP_Input_PVT‖

worksheet

171

172

b. Simulate standard PVT properties from compositional data

i. In ―PVT_Input_Pres‖ worksheet, input reservoir and surface separators

pressure and temperature conditions

ii. In ―PVT_Input_Comp‖ worksheet, input composition (mole fraction)

and properties of pure components

iii. In ―PVT_Input_BiCo‖ worksheet, input binary interaction coefficient

iv. In ―PVT_Input_Si‖ worksheet, input volume-translate coefficient data

v. In ―PVT_Output_Prop‖ worksheet, click ―Calculate Black Oil PVT

Properties‖ button

vi. In ―PVT_Output_Envelope‖ worksheet, click ―Create Phase Envelope‖

button (optional)

vii. In ―FDP_Input_PVT‖ worksheet, click ―Import from PVT Calculation‖

button

173

174

175

176

3. Perform field performance analysis

a. Run field performance prediction and economic analysis

i. In ―FDP_Input_Tank‖ worksheet, input reservoir data

ii. In ―FDP_Input_RelPerm‖ worksheet, input relative permeability data

iii. In ―FDP_Input_TPR‖ worksheet, input tubing data

iv. In ―FDP_Economic‖ worksheet, input economic data

v. In ―FDP_Input_Main‖ worksheet, input target recovery factor at end of

plateau and total number of wells required for field development

vi. Click ―Run Performance Prediction and Economic Analysis‖ button

vii. Obtain economic results from ―FDP_Economic‖ worksheet

viii. Obtain field performance data from ― FPD_Output_Perf‖ worksheet

ix. Obtain reservoir performance data from ―FDP_Output_SGCT‖

worksheet

x. Obtain decline trend analysis data from ―DCA‖ worksheet

xi. Obtain graphical display of flow rates, pressures and cumulative

production from ―QgGp‖, QoNp‖, and QwPrPwfPwh‖ worksheets

177

178

179

180

181

182

b. Run economic analysis from latest field performance data

i. In ―FDP_Economic‖ worksheet, input economic data

ii. In ―FDP_Input_Main‖ worksheet, click ―Run Economic Analysis from

Latest Performance Data‖ button

iii. Obtain economic results from ―FDP_Economic‖ worksheet

183

184

4. Perform field optimization

a. Run optimization with new performance prediction and economic analysis

i. Repeat 3.-a.-i ) to 3.-a.-iv)

ii. In ―FDP_Input_Main‖ worksheet, input target recovery factor at end of

plateau and total number of wells required for field development

iii. Click ―Optimization 1: Run Performance Prediction and Economic

Analysis‖ button

iv. Obtain field optimization results from ―Optimization‖ worksheet

185

b. Run optimization with new economic analysis only (utilize latest performance

prediction data)

i. In ―FDP_Economic‖ worksheet, input economic data

ii. In ―FDP_Input_Main‖ worksheet, click ―Optimization 2: Run Economic

Analysis only‖ button

iii. Obtain field optimization results from ―Optimization‖ worksheet

186

187