You are on page 1of 470

Petroleum Engineering

Reza Azin
Amin Izadpanahi Editors

Fundamentals and
Practical Aspects of
Gas Injection
Petroleum Engineering

Editor-in-Chief
Gbenga Oluyemi, Robert Gordon University, Aberdeen, Aberdeenshire, UK

Series Editors
Amirmasoud Kalantari-Dahaghi, Department of Petroleum Engineering, West
Virginia University, Morgantown, WV, USA
Alireza Shahkarami, Department of Engineering, Saint Francis University, Loretto,
PA, USA
Martin Fernø, Department of Physics and Technology, University of Bergen,
Bergen, Norway
The Springer series in Petroleum Engineering promotes and expedites the dissem-
ination of new research results and tutorial views in the field of exploration
and production. The series contains monographs, lecture notes, and edited volumes.
The subject focus is on upstream petroleum engineering, and coverage extends to
all theoretical and applied aspects of the field. Material on traditional drilling and
more modern methods such as fracking is of interest, as are topics including but not
limited to:
• Exploration
• Formation evaluation (well logging)
• Drilling
• Economics
• Reservoir simulation
• Reservoir engineering
• Well engineering
• Artificial lift systems
• Facilities engineering

Contributions to the series can be made by submitting a proposal to the


responsible publisher, Anthony Doyle at anthony.doyle@springer.com or the
Academic Series Editor, Dr. Gbenga Oluyemi g.f.oluyemi@rgu.ac.uk.

More information about this series at http://www.springer.com/series/15095


Reza Azin · Amin Izadpanahi
Editors

Fundamentals and Practical


Aspects of Gas Injection
Editors
Reza Azin Amin Izadpanahi
Petroleum Engineering Oil and Gas Research Center
Persian Gulf University Persian Gulf University
Bushehr, Iran Bushehr, Iran

ISSN 2366-2646 ISSN 2366-2654 (electronic)


Petroleum Engineering
ISBN 978-3-030-77199-7 ISBN 978-3-030-77200-0 (eBook)
https://doi.org/10.1007/978-3-030-77200-0

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Switzerland AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Gas injection has been used as a means of pressure maintenance in the secondary oil
recovery and in the tertiary oil recovery as Enhanced Oil Recovery (EOR) techniques.
The concept and operation of gas injection started nearly two decades after industrial
oil production in the nineteenth century when the operators faced with the pressure
decline in the reservoir and found the produced gas as an abundant source to reinject
and overcome the problem. Many field cases used various injecting gases, ranging
from non-hydrocarbon to hydrocarbon, lean to enriched, or immiscible to miscible.
The operating conditions vary from one case to another, depending on the geological
conditions of the reservoir, its rock and fluid properties, and the availability of gas
sources to inject. The objective of gas injection may also be different from one project
to another. This book reviews some of the applications of gas injection and combines
the new features as well as some key challenges associated with gas injection.
The book is organized into ten chapters. Chapter 1 provides an overview of the
hydrocarbon resources worldwide as well as a historical review of the gas injec-
tion operation. In Chap. 2, some of the key issues related to Pressure–Volume–
Temperature (PVT) are reviewed. The phase diagram of injection gas, reservoir oil,
and their mixtures are presented and discussed. Also, important PVT experiments,
including CCE, DL, CVD, flash separation, and swelling test are discussed and
supported by worked examples. Chapter 3 deals with basic concepts of oil and
gas flow in porous media. In this chapter, single- and two-phase flow concepts
are reviewed. Different flow equations, including viscous, pre-Darcy, Darcy, and
Brinkman flow are introduced and the boundary between these flow regimes is
presented. The well-known diffusivity equation which is a combination of conti-
nuity equation and Darcy velocity is also introduced and discussed. The relative
permeability as an important concept in two-phase flow through porous media is
described and the forces affecting oil flow in porous media are introduced as well
as different dimensionless numbers which reflect the relative importance of forces
during oil flow. Chapter 4 focuses on Underground Gas Storage (UGS) as a special
gas injection applied for peak shaving and balancing the supply and demand of
natural gas. In Chap. 5, gas injection into the gas cap known as pressure maintenance
is introduced. The concept of pressure maintenance, historical approach, mechanistic

v
vi Preface

modeling, and reservoir material balance are discussed, followed by pressure main-
tenance models and mechanisms in fractured reservoirs. Chapter 6 deals with gas
recycling and addresses conceptual, economical, historical, and operational aspects
of the gas recycling operation. Chapter 7 focuses on the design of surface facilities
for gas injection. In this chapter, conceptual design, design considerations and basis,
and the general layout of a gas injection facility are described. Then, elements of a
surface facility are introduced and described in detail. Chapter 8 presents methods
for predicting the water content of gases, mostly based on equations of state and
rigorous thermodynamic models. Different methods of predicting the water content
of acid gas systems are evaluated and the water content diagrams compatible with
the experimental data for pure CO2 , H2 S, CH4 , and their mixture are presented.
Also, the hydrocarbon and non-hydrocarbon gases solubility in water, calculation
of gas solubility in water using Henry’s constant law and the effect of salinity on
gas solubility are presented in this chapter. The structure of Chap. 8 is based on
engineering procedures and worked examples and may be used as an engineering
guide to predict the water content of gases and gas solubility in brine. Chapter 9
focuses on the most important challenges including the corrosion in different types
of gas injection, gravity override, mobility control, cap rock integrity, trapping, and
Health, Safety and Environment (HSE) of gas injection. These concepts must take
into account before, during, and after a gas injection project to increase the effi-
ciency of the project. Chapter 10 deals with capillary trapping in porous media.
The concepts, types, models, and correlations for phase trapping are presented and
discussed in this chapter. References are added at the end of each chapter. Also,
some chapters contain useful information as appendix. In preparing the contents of
this book, more than one thousand references were reviewed comprehensively. Also,
many chapters include worked examples to make a clear content and field application

Bushehr, Iran Reza Azin


March 2021 Amin Izadpanahi
Acknowledgements

A number of scientists, professors, and industrial experts contributed to the chapters


of this book. These include Prof. Mahmood Moshfeghian, Prof. Shahriar Osfouri, Dr.
Ahmad Jamili, Dr. Ali Ranjbar, Dr. Alireza Shahkarami, Dr. Mohamad Mohamadi-
Baghmolaei, Ahmad Banafi, Amin Izadpanahi, Fatemeh Kazemi, Parviz Zahedi-
zadeh, and Pooya Aghaee Shabankareh. Also, the authors visited several oil and gas
injection and production facilities and had technical meetings with industrial experts
to get better insight into details of gas injection operations. I’d like to acknowl-
edge Reza Heidari, Mohammad Abdali, and Dr. Hamid Khedri and appreciate their
technical support and valuable suggestions.

Reza Azin
Professor of Petroleum Engineering
reza.azin@pgu.ac.ir

vii
Contents

1 Introduction to Gas Injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


Reza Azin, Amin Izadpanahi, and Alireza Shahkarami
2 PVT of Gas Injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Reza Azin, Amin Izadpanahi, and Shahriar Osfouri
3 Basics of Oil and Gas Flow in Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . 73
Reza Azin, Amin Izadpanahi, and Parviz Zahedizadeh
4 Gas Injection for Underground Gas Storage (UGS) . . . . . . . . . . . . . . . 143
Reza Azin and Amin Izadpanahi
5 Gas Injection for Pressure Maintenance in Fractured
Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Ahmad Jamili, Amin Izadpanahi, Pooya Aghaee Shabankareh,
and Reza Azin
6 Gas Recycling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
Reza Azin, Amin Izadpanahi, and Mohamad Mohamadi-Baghmolaei
7 Design of Subsurface and Surface Facilities for Gas Injection . . . . . . 319
Reza Azin and Ahmad Banafi
8 Water-Hydrocarbons System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Amin Izadpanahi and Reza Azin
9 Challenges of Gas Injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
Reza Azin, Amin Izadpanahi, and Ali Ranjbar
10 Capillary Phase Trapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
Fatemeh Kazemi, Reza Azin, and Shahriar Osfouri

ix
Chapter 1
Introduction to Gas Injection

Reza Azin, Amin Izadpanahi, and Alireza Shahkarami

Abstract This chapter describes the different methods and the application of gas
injection in the oil and gas industry and environmental purposes. For this purpose, first
the worldwide hydrocarbon distribution is studied which categorized as conventional
and unconventional resources. Also, the graphical distribution is reported using the
latest statistics of conventional and unconventional hydrocarbon resources. Second,
the steps of recovery of oil reservoirs are studied which consist of primary, secondary
and tertiary oil recovery. Gas injection is performed for improvement of oil recovery
in the secondary and tertiary oil recovery. Environmental protection and reduction
of greenhouse gases is another goal of gas injection. Third, various relevant aspects
including, historical evolution, screening criteria, sources of gas and economic of
gas injection are investigated.

1.1 Introduction

Rapid economic growth and population increment have led to a severe increase in
energy demand in all around the world. According to the BP statistical review of
world energy in 2016, fossil fuels among different natural energy resources, have
the largest contribution (nearly 80%) in supplying the world’s energy needs. Also, it
has been estimated that fossil fuels will remain the predominant sources of energy
by 2035 [1]. Nonetheless, these most widely used energy sources have numerous

R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
A. Shahkarami
Saint Francis University, Loretto, PA, USA
e-mail: ashahkarami@francis.edu

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 1


R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_1
2 R. Azin et al.

negative aspects such as the economic dependency of some regions and countries
on fossil fuels, their reserves restriction and most importantly, environmental effects
[2].
With regard to the global energy consumption in the last three decades, oil and
gas are the most important hydrocarbons and fossil resources [1]. These resources
are mainly classified into two categories, named conventional and unconventional.
Global distribution of these two groups happens at a ratio of about 1:4 [3]. Various
definitions have been appointed for conventional and unconventional reservoirs by
numerous experts. Nevertheless, the most prevalent description of conventional
resources is supreme quality and high permeability reservoir which may be with-
drawn economical volumes of hydrocarbons without needing substantial excitation
and only by drilling a vertical hole and perforating the productive interval [4]. As
reported by the international energy agency (IEA), these hydrocarbon reservoirs
are largely concentrated in the Middle East and Russia. Global proven reserves of
conventional oil and gas are prognosticated about 1.3 trillion barrels and 220 trillion
cubic meters, respectively [5]. Any source of oil and gas which has a low quality
reservoir due to the low permeability or high oil viscosity and needs production
technologies remarkably different from those used to produce from conventional
reservoirs is commonly illustrated as unconventional [6]. Canada, Venezuela and
Asia Pacific are the main regions with these types of hydrocarbon resources. Oil
shale, oil sands (heavy oils, bitumen and tar deposits), light tight oil, tight gas, shale
gas, coal-bed methane (CBM) and methane hydrates are the categories of unconven-
tional resources as defined by IEA [5]. Estimated unconventional oil proven reserves
are around 400 billion barrels in the world. Due to the high heterogeneity of rock
formations, assessment of proven reserves is very arduous in unconventional gas
resources, and its exact value is uncertain [5].
As mentioned, another important drawback of a sharp increase in fossil fuel
consumptions is CO2 and other greenhouse gases (GHGs) emissions. The increase
of GHGs in the atmosphere, has led to environmental disruption through global
warming and climate changes [7]. Therefore, environmental protection with regard
to the fossil fuels domination on worldwide energy supply, has become a global
challenge in petroleum industry.
Gas injection has been suggested and employed as an efficacious approach almost
since 1930 to extend hydrocarbon production in different stages of exploitation,
called secondary and tertiary recovery operations [8–11]. Also, in the case of CO2 or
acid gas injection, this procedure has been considered as useful tools to reduce the
concentration of atmospheric GHGs and its destructive effects [12]. Oil reservoirs
due to the very low primary recovery factor, around 20–40% for conventional and
up to 15% for unconventional, in comparison with gas reservoirs were aimed at
most gas injection projects in all around the world [13]. This chapter provides an
overview of worldwide hydrocarbon resource distribution, followed by a review on
the state-of-the art of gas injection technologies into oil reservoirs as secondary and
tertiary recovery processes. Various relevant aspects, including historical evolution,
the goal of gas injection, sources of gas for injection, challenges and problems of gas
1 Introduction to Gas Injection 3

injection will be presented, and some key research findings and the gaps in current
understanding will be highlighted.

1.2 Worldwide Distribution of Hydrocarbon Reservoirs

Hydrocarbon reservoirs generally categorized as conventional and unconventional


reservoirs. This grouping is defined based on the required technology for extracting
the hydrocarbon and geological reservoir conditions.
Unconventional oil and gas extraction needs new and costly technologies
comparing to conventional ones. Conventional oil is identified as the oil produced
by the reservoir’s energy and pressure maintenance methods (primary and secondary
recovery). In contrast, unconventional oil cannot produce without the enhanced oil
recovery methods. Extra heavy oil, natural bitumen, tar sands, and shale oil are
categorized as unconventional oil [14].
Conventional natural gas is defined as the gas produced by reservoir pressure, but
the production of unconventional gas requires complicated technologies. Tight gas,
coal-bed methane, and gas hydrates are known as unconventional gas resources [14].
Figure 1.1 shows the resource triangle of conventional and unconventional resources
[6].
In this section, the worldwide distribution of each resource is explained.

Fig. 1.1 Resource triangle which comprises the volume of conventional and unconventional
resources [6]
4 R. Azin et al.

1.2.1 Conventional

There are some annual reports that contain the amounts of energy resources which
published by the British Petroleum (BP) company [15, 16]. In this chapter, the
amounts of crude oil and natural gas are reported from these annual reports of 2010
and 2018. As it is obvious, these amounts are changed, and the share percent of each
region increased or decreased.
Crude Oil. The Middle East, Latin America and North America have the most
conventional oil resources. Table 1.1 shows the proven oil reserves in different regions
of the world in 2010 and 2018.
Natural Gas. Europe and Eurasia and the Middle East have the most amount
of natural gas reserves which reported in Table 1.2. In this table, the amount of
conventional gas reserves in 2010 and 2018 has been reported.

Table 1.1 Distribution of crude oil reserves all over the world [15, 16]
Area Proven oil reserves Percentage (%) Proven oil reserves Percentage (%)
2010 (109 bbl) 2018 (109 bbl)
North America 73.7 5.5 226.1 13.3
Latin America 198.9 14.9 330.1 19.5
Europe and Eurasia 136.9 10.3 158.3 9.3
Middle East 754.2 56.6 807.3 47.6
Asia 42.2 3.2 48 2.8
Africa 127.7 9.6 126.5 7.5
Total (109 bbl) 1333.6 100 1696.3 100

Table 1.2 Distribution of natural gas reserves all over the world [15, 16]
Area Proven natural gas Percentage (%) Proven natural gas Percentage (%)
reserves 2010 reserves 2018
(1012 m3 ) (1012 m3 )
North America 9.16 4.9 10.8 5.6
Latin America 8.06 4.3 8.2 4.2
Europe and Eurasia 63.09 33.7 62.2 32.1
Middle East 76.18 40.6 79.1 40.9
Asia 16.24 8.7 19.3 10
Africa 14.76 7.9 13.8 7.1
Total (1012 m3 ) 187.49 100 193.4 100
1 Introduction to Gas Injection 5

1.2.2 Unconventional

The unconventional gas resources include coal-bed methane, tight gas, aquifer gas,
and gas hydrates. Oil sands, heavy oil, natural bitumen and shale oil are categorized
as the unconventional oil resources [14].
Oil Sands. Oil sands (also known as natural bitumen and tar sands) are mixtures of
water, sands, bitumen and clay. The density of oil sands is greater than 1 gr/cm3 (less
than 10° API) and viscosity is greater than 10,000 cp. The methods of producing oil
sands include surface mining, underground mining or thermal methods depending
on the reservoir depth. Cyclic steam stimulation (CSS) and steam-assisted gravity
drainage (SAGD) are applied to reduce oil viscosity and make the fluid mobile. In
CSS, the viscosity is reduced by injecting the steam and held the steam in the reservoir
for a specified time. In SAGD, two horizontal wells are drilled from a vertical well.
The steam is injected from the upper well and the oil is produced from the lower
well [14].
Heavy Oil. The definition and classification of heavy oil differ from one country
to other. In countries having large resources of conventional crude oil, classification
of heavy oil refers to oils with 20–25 API. However, in some regions, heavy oil refers
to oils below 20 API and viscosity lower than that (200 cp). The conventional, as
well as thermal methods, may be applied to extract heavy oil from reservoirs.
Shale Oil. A calcareous mudstone which contains kerogen is known as shale oil.
The methods of producing shale oil are similar to oil sands. About 35–245 Liters of
oil can be extracted from 1-ton shale oil rock [14]. Table 1.3 shows proven reserves
of unconventional oil in the world [17, 18]. Also, Fig. 1.2 shows the distribution of
unconventional oil in the world.
Shale Gas. The gas trapped in the fine-grained sedimentary rock (shale) is known
as shale gas. Normally, hydraulic fracturing is required for the production of this
kind of gas.

Table 1.3 Distribution of unconventional oil around the world [17, 18]
Region Oil sands (109 Heavy oil (109 Shale oil (109 bbl) Total (109 bbl)
bbl) bbl)
North 530.9 35.3 80 646.2
America
Latin America 0.1 265.7 59.7 325.5
Europe and 33.9 18.3 90.1 142.3
Eurasia
Middle East 0 78.2 42.9 121.1
Asia 42.8 29.6 73.9 146.3
Africa 43 7.2 0.1 50.3
Total 650.7 434.3 346.7 1431.7
(109 bbl)
6 R. Azin et al.

Fig. 1.2 Distribution of unconventional oil in different regions of the world (references are the
same as Table 1.3)

Tight Gas. The gas reservoirs with the permeability less than 0.1 millidarcy
are known as tight gas reservoirs. Low permeability sandstone or limestone is the
main rock in the tight gas reservoirs [19]. The main method of producing tight gas
is hydraulic fracturing technique that entails injecting high pressure mixtures of
water, proppants, and chemical to crack the reservoir rock and consequently enhance
reservoir permeability [14].
Coal-bed Methane. The coal seams contain rich methane and often high propor-
tions of nitrogen and carbon dioxide. These kinds of gas resources are known as
coal-bed methane, where coal is the source rock of methane. The method of producing
coal-bed methane is drilling a well in the coal seam and creating artificial fractures
and filling them with a sand-water mixture. Table 1.4 shows the amount of uncon-
ventional gas resources all around the world. Also, Fig. 1.3 shows the distribution of
unconventional gas in the world.

Table 1.4 Unconventional gas resources all around the world [19]
Area Tight gas (1012 m3 ) Coalbed methane Shale gas (1012 m3 ) Total
(1012 m3 )
North America 22.3 6.3 42.7 71.3
Latin America 3.7 0.2 35.6 39.5
Europe and Eurasia 4.9 1.7 51.5 58.1
Middle East 2.3 0 15.5 17.8
Asia 17 18.8 43.3 79.1
Africa 2.3 0.9 29.3 32.5
Total (1012 m3 ) 52.5 27.9 217.9 298.3
1 Introduction to Gas Injection 7

Fig. 1.3 Distribution of unconventional gas in different regions of the world (references are the
same as Table 1.4)

Figures 1.4 and 1.5 compare the conventional and unconventional oil and gas
resources worldwide, respectively.

Fig. 1.4 Comparison of distribution of conventional and unconventional oil in different regions of
the world (references are the same as Tables 1.1 and 1.3)
8 R. Azin et al.

Fig. 1.5 Comparison of the distribution of conventional and unconventional gas in different regions
of the world (references are the same as Tables 1.2 and 1.4)

1.3 Historical Evolution of Gas Injection

Gas injection is the oldest method of injection fluid into the hydrocarbon reser-
voirs, with various applications in oil and gas industry and it has been done for
different purposes such as pressure maintenance, enhanced oil recovery (EOR),
underground gas storage (UGS), carbon capture and sequestration (CCS), gas recy-
cling, and etc. Pressure maintenance, EOR and gas recycling have been done for
improving the ultimate recovery from oil and gas reservoirs. These methods can
result in recovery improvement by affecting different parameters such as reservoir
pressure, oil viscosity, miscibility conditions, capillary forces, sweep efficiency, and
etc. Underground gas storage is a method to store the hydrocarbon gases in natural
reservoirs to meet the energy demands during high consumption seasons. Carbon
capture and sequestration is a necessary process for decreasing the effects of green-
house gases and preventing global warming. In this section, a brief history of each
method is expressed.
The concept of gas injection was first suggested as early as 1864 [20]. In 1903,
Dunn expressed the results of gas injection into a well. Then, studies continued on gas
injection and the first re-pressuring project was carried on in Ohio by air injection into
depleted and partially depleted wells in 1913. The results of this project demonstrated
that the production from these wells greatly increased after gas injection. After this
project, pressure maintenance was widely used in different fields worldwide as an
improving oil recovery method [21].
Underground gas storage started with an experiment as early as 1915 in Ontario,
Canada. Gas was injected into several reconditioned old wells. Gas production started
from that field in the following winter to meet peak requirements. This experiment
1 Introduction to Gas Injection 9

repeated the next year (1916) in the near Buffalo, New York. After the success of
these two projects, underground gas storage became a common method for storing
hydrocarbon gases in low consumption seasons and producing these gases in high
consumption seasons [22].
Gas recycling projects began in 1941. In one case, gas recycling was done to
improve the recovery from wet-gas reserves. In another study, gas recycling was
proposed as a production method based on 8 years of field experiments [23, 24].
The growth of EOR processes has been continued since the end of world war
II. In that time, operators who had faced with declining reservoirs production after
primary and secondary recovery found out that large quantities of oil stayed in their
reservoirs. Decreasing production from reservoirs, increasing oil consumption, and
decreasing success of new reservoir discoveries motivated the research and field
activities involving EOR processes to respond in specific the oil boycott of 1973
and the subsequent energy crisis. Economic efficiency is one of the most important
parameters for planning an EOR project, in other words, the costs of producing oil
able to compete economically with the cost of EOR processes [25]. In the 1950s,
miscible flooding was shown as one of the most favorable methods for enhancing
ultimate oil recovery from reservoirs [26–32]. The significant viscosity reduction
effect of carbon dioxide injection was reported as early as 1963. It has been shown
that the viscosity reduction of heavy oil using carbon dioxide could be up to 98%
[33]. In the 1970s, studies were conducted in the field of CO2 viscosity reduction
and CO2 huff ‘n’ puff process [34, 35]. The first cyclic stimulation with CO2 was
done in 1981 in Canada [34].
Application of carbon capture to mitigate climate change was first discussed in
1982 by focusing on control CO2 production from power plants [36, 37]. There are
other alternatives accessible in the case of CO2 storage. The first suggested option
for carbon storage was as early as 1977. In this proposal, the injection of carbon
dioxide into the ocean was expressed. In this method, CO2 is transferred into deep
water where it would stay for many years [38]. The other options for storing carbon
are geological formations, oil and gas reservoir for EOR purposes, saline formations
and coal seams [39–42]. The world’s first CO2 storage plant began to operate in
1996. In this plant, CO2 injected into a deep saline formation under the North Sea
[43]. Figure 1.6 shows the historical timeline of gas injection methods.

1.4 The Need for Gas Injection

Figure 1.7 shows the conceptual trend of production rate and cumulative produc-
tion with time. As shown in this figure, throughout the lifecycle of an oil field,
various stages of production are encountered. Primary crude oil production frequently
proceeds with natural energies of a reservoir to let the hydrocarbons flow to the
surface. In this stage, production rate experiences a buildup phase followed by plateau
rate. The primary production mechanisms include rock and fluid expansion, solution
gas drive, gravity segregation, gas cap drive, natural water influx, or a combination
10 R. Azin et al.

Fig. 1.6 Historical evolution of gas injection

of two or more energies. Contribution of each of these driving forces in pushing


the oil towards and out of producer well is different from one formation to another
and depends on a number of factors such as geological structure, petrophysical char-
acteristics, physical and thermodynamic properties of reservoir fluids, and etc. By
declining the reservoir pressure, the natural drive energies of reservoir decrease and
external forces are required to keep the reservoir production.
Studies indicate that solution gas drive exists in the most oil reservoirs and can
recover less than 25% of original oil in place (OOIP) for conventional reservoirs
[44, 45]. Also, natural water drives exist almost in one-third of world reservoirs. In
the most optimistic case, a strongly water driven conventional reservoir with high
permeability, up to 50% of OOIP can be extracted [45, 46]. For gas-cap expansion
and gravity drainage, the recovery is only around 5–30% [46]. Usually, the primary
recovery of unconventional reservoirs is much lower than conventional ones. Studies
on shale formations and extremely low-permeable oil reservoirs show that the primary
recovery is below 10% [47–49]. Recovery factors of heavy oil reservoirs (<22.3°
API) are reported in a range of 10–15%. It should be mentioned that highly viscous
hydrocarbon resources like bitumen and tar sands may be incapable of being produced
1 Introduction to Gas Injection 11

Fig. 1.7 Schematic representation of a petroleum reservoir lifecycle

by primary reservoir drives and require sophisticated technologies to produce from


the early life of reservoir [45]. On the other side, natural depletion in oil reservoirs
alone yields low recovery factors, as reported by IEA averagely 5–15% irrespective
of the reservoir type [5]. As a result of this impediment, large and considerable oil
volumes remain underground at the end of the primary recovery stage.
The upstream oil and gas experts and scientists attempt to find efficient technolo-
gies to recover more hydrocarbons in place and extend the lifecycle of oil fields.
The use of these technologies can make an important contribution to secure long-
term oil supply. At the end of primary depletion, the oil production rate can increase
through injection of gas into reservoir, as known as secondary recovery stage. In
this stage, immiscible gases are injected into a reservoir gas cap to supply reservoir
energy, also known as pressure maintenance. This operation provides the reservoir
with sufficient energy to continue oil production while storing the gas into gas cap.
By increasing the reservoir pressure, the produced gas oil ratio (GOR) decreases
and the oil rate increases. As a result, the ultimate recovery from the reservoir will
be improved [50]. In the reservoirs where gravity drainage is the main production
mechanism, pressure maintenance by gas injection is found a successful recovery
technique. Ultimate recovery could be reached about 50% of the initial oil in place
by pressure maintenance using gas injection [51, 52]. Literature indicates that gas
injection is superior over water injection in low-permeability reservoirs (such as
fractured shales), swelling clay and steeply-dipping ones [44, 53].
12 R. Azin et al.

Studies indicate that around 15–20% and 5–8% of OOIP may be produced by
secondary recovery methods, i.e. water or gas injection, from conventional and
unconventional oil reservoirs, respectively [46, 54]. Despite utilizing these IOR
processes, still remarkable oil volumes remain unrecovered in reservoirs. Further
oil recovery may be obtained through enhanced oil recovery (EOR) technologies,
also known as tertiary oil recovery. According to the global energy consumption
report by IEA, a 1% increase in average recovery factor can add more than 6% to the
global proven hydrocarbon reserves [55]. Enhanced oil recovery (EOR) processes
are long-term solutions to improve oil production [56, 57] applied to the mature,
depleted reservoir to increase oil recovery.
The gas injection applied for EOR may be classified either as an immiscible or
miscible gas injection. The immiscible gas sweeps the reservoir oil without dissolving
into it, thus creating a more or less sharp front between phases during displacement.
On the other side, the miscible gas dissolves into reservoir oil while displacing it. The
injected gas may dissolve into reservoir oil and at the same time, reservoir oil may
vaporize into injection gas. The degree of miscibility, i.e. dissolution and vaporization
depends on the composition of injection and reservoir fluids, reservoir temperature
and pressure, and the time of contact which affects the approach or deviation from
thermodynamic equilibrium. Miscibility can be formed in the first contact, known as
first-contact miscibility (FCM), or through multiple contacts between phases, known
as multiple-contact miscibility (MCM).
The immiscible gas injection may be combined with water flooding, so called
water alternating gas injection, to reduce the segregation caused by density differ-
ences between oil and gas. For heavy oil reservoirs with high initial viscosity, the
injection gas may act as a “solvent” and reduce the oil viscosity substantially and
mobilize it in the vapor extraction (VAPEX) process. Details of this process are
described by Azin et al. [58–63]. The concept of solvent implies a phase that dissolves
a solute. In the case of VAPEX, the gas that dissolves into heavy oil is considered
as solvent, while the heavy oil shows a little chance to vaporize into injected gas.
Therefore, the miscibility is limited to the dissolution of injected gas into reservoir
oil.
Despite all of its advantages, the gas injection has some disadvantages in some
reservoirs. For example in high GOR reservoirs, gas injection is not recommended
due to reduction of final hydrocarbon recovery. Izadpanahi et al. studied an Iranian
fractured carbonate reservoir with high producing GOR. They studied six scenarios
for development of the mentioned field including natural depletion, water injection,
gas injection, infill drilling and simultaneous water and gas injection. Their results
showed that gas injection is not suitable in the studied reservoir. This scenario has
the lowest final hydrocarbon recovery among other scenarios [64, 65].
Gas injection has also found its crucial role in environmental protection through
carbon capture and sequestration (CCS) of greenhouse gases (GHG). The world-
wide concerns about global warming and the rise of CCS projects to mitigate the
environmental challenges of GHG have led to a new paradigm of gas injection
into underground formations, including depleted oil and gas reservoirs and saline
aquifers. Details of CCS projects, including planning, transportation and injection
1 Introduction to Gas Injection 13

Table 1.5 CCS components and technologies [66]


CCS component CCS technology
Capture Post-combustion
Pre-combustion
Oxyfuel combustion
Industrial separation (natural gas processing, ammonia production)
Transportation Pipeline
Shipping
Geological storage Enhanced oil recovery (EOR)
Gas or oil field
Saline formation
Enhanced coal bed methane recovery (ECBM)
Ocean storage Direct injection (dissolution type)
Direct injection (lake type)
Mineral carbonation Natural silicate minerals
Waste materials
Industrial uses of CO2 –

may be found in the international reports such as intergovernmental panel on climate


change (IPCC) [66]. Also, an updated data bank and information of CCS projects
may be found in the Global CCS Institute website (https://www.globalccsinstitute.
com/) and carbon capture and sequestration association (www.ccsassociation.org).
Table 1.5 summarizes the CCS steps and technologies [66].
Another area that depends on gas injection covers the projects which inject and
withdraw natural gas in successive cycles to mitigate the gap between supply and
demand of natural gas during the year. Underground gas storage (UGS) has been
used for more than one century since 1915 and is found a promising technique to
supply a constant energy stream. The UGS may be operated by injecting natural gas
into depleted oil and gas reservoir, saline aquifers, and salt caverns. In this approach,
any type of gas, e.g. dry natural gas, sour gas containing H2 S and CO2 , etc. may be
stored [67–70].
One of the most common processes for maintaining reservoir pressure above the
bubble or dew point pressure is gas recycling [71]. Gas recycling is a special type of
pressure maintenance in which the injected gas is produced from the same reservoir
and reinjected into gas cap after separation from produced oil. Depending on the
development strategy and downstream plans, the associated gas may be partially
or totally reinjected into the reservoir. Also, gas recycling can be operated in gas
condensate reservoirs to vaporize the condensate accumulated around the wellbore.
14 R. Azin et al.

1.5 Screening Criteria for Gas Injection

Table 1.6 summarizes the approximate ranges of the fluid and reservoir properties for
good projects. Taber et al. reported the approximate ranges of criteria for miscible
flood projects. For example, nitrogen and flue gas injection are suitable for oil lighter
than 35° API and the reservoir with a minimum depth of 6000 ft. This kind of gas
becomes miscible with reservoir fluid in the reservoirs with high pressure due to high
minimum miscibility pressure. The good criteria for other miscible floods are shown
in Table 1.6.
The criteria for the huff-n-puff project are reported based on 31 projects that
collected from the literature. Most of these projects are in the United States of
America. The porosity ranges in these projects vary between 0.1 and 0.33 but these
fields have a wide range of permeability. The other important characteristics are
reported in Table 1.6.
The ranges of porosity in the reservoirs under pressure maintenance operations
using gas injection are 0.0758–0.336 and the permeability in the ranges of 40–5000
md. Pressure maintenance can be carried out in a wide range of reservoir pressure
(1210–5029 psi). This information are extracted from 21 pressure maintenance field
projects.
The screening criteria for gas recycling is achieved by collecting the information
from field projects and simulation studies on gas recycling scheme. The literature
consists of both gas condensate and oil reservoirs under gas recycling operation. So,
the fluid properties can’t select as screening criteria parameters. The porosity ranges
are reported between 0.06–0.24 and the permeability are reported in the ranges of
22–2990 md. Other information can be seen in Table 1.6.

Table 1.6 Screening criteria for different gas injection methods


Gas injection methods Fluid properties Reservoir properties References
Gravity Viscosity Thickness Depth (ft)
(API) (cp) (ft)
Miscible Nitrogen and >35 <0.4 Thin unless >6000 [72, 73]
flood flue gases dipping
Hydrocarbon >23 <3 Thin unless >4000
dipping
CO2 >22 <10 Thin unless >2500
dipping
Huff n puff >0.775 <44.7 7–222 2365–13,140 [74–79]
Pressure maintenance >0.026 18.9–42 43–1800 2000–11,500 [51, 52, 80–86]
Gas recycling NM NM 21–1080 5855–13,205 [24, 87–98]
NM = Not mention
1 Introduction to Gas Injection 15

1.6 Sources of Gas for Gas Injection

The basic requirement for a gas injection operation is the availability of a certain
source of gas. Generally, either hydrocarbon or non-hydrocarbon gases may be used
for injection. Some sources like CO2 as the main greenhouse gas with increasing
emission trends have been found an attractive source for injection.
Also, N2 is another abundant source of gas and has been considered as a suitable
gas for pressure maintenance projects in both oil and gas condensate reservoirs.
Nitrogen is a safe, noncorrosive and environment friendly gas which can be separated
from air. However, it raises the saturation pressure of reservoir fluid; moreover, it
can’t dissolve in heavy hydrocarbon components and has a high minimum miscibility
pressure. Therefore, N2 injection as an enhanced oil recovery process can be applied
to high pressure, deep and light oil reservoirs [99–103].
The mixture of N2 and CO2 which is known as flue gas can be used as an alternative
to pure CO2 and pure N2 for gas injection. Injection of this mixture has some advan-
tages such as eliminating the pure gas separation costs, postponing the gas break-
through compared to pure N2 injection, low compressibility and occupying more
space in the reservoir compared to other gases and enhance the production because
of CO2 displacement [104–106]. The mass transfer mechanism is also included in
the flue gas injection which is in favor of valuable oil components recovery [107]. It
was shown in the literature that the flooding efficiency is increasing by increase the
CO2 fraction in the flue gas due to CO2 ability to crude oil swelling [108].
Another gas injection source is the hydrocarbon stream supplied either by asso-
ciate gas and used for gas recycling or from independent gas reservoirs. The hydro-
carbon gas may be used before or after treatment in gas plant. In other words,
the injection gas may contain hydrocarbons, mostly methane, and non-hydrocarbon
components like CO2 , H2 S, N2 , noble gases, and etc. It was shown that hydrocarbon
gas injection works better than N2 and CO2 injection because of the adaptability
and miscibility between the injected gas and reservoir fluids. The disadvantage of
this method is that a large amount of natural gas will be back into the reservoir and
this amount is unusable until the gas recycling project is stopped [106]. Therefore,
hydrocarbon gas injection is attractive in areas with abundant sources of natural gas.
Also, hydrocarbon gas injection has been used in gas condensate reservoirs to reduce
the amount of condensate near wellbores, maintain the reservoir pressure above the
saturation pressure and decrease the oversupply operation in the low market seasons
[109].
The mixture of CO2 and H2 S, which is called Acid gas, could be considered as
an injection fluid. This mixture can be released with the upstream production and
the decrement gases from natural gas and petroleum processing. Acid gas has many
irreparable effects on human health because of that in order to dispose, acid gas
injection must be taken into account [110–112].
16 R. Azin et al.

1.7 Economics of Gas Injection

For developing a gas injection process an economic evaluation must take into account
for measuring the feasibility of the method. In this section, a review of the economic
evaluation of different modes of gas injection was done.

1.7.1 Pressure Maintenance

For planning a pressure maintenance project an economic analysis should be done.


The expenditures of a pressure maintenance project include, well operation, plant
and lines costs. The well operation costs are different for a flowing well with a
pumping well. The costs of operating a flowing well will be lower than that of a
pumping well and the average cost of a flowing well will increase more slowly than
that of a pumping well. Also, replacements and workovers must be considered for
well operation costs.
In a pressure maintenance project, the operation of some plants and lines are
necessary such as, compressor plant, injection lines, de-sulphurizer, and dehydrator.
The costs of labor, supervision, repairs and materials must be added to the costs of
building and operating the mentioned plants and lines [113].

1.7.2 CCS

Some factors influence the costs of CCS projects such as fuel prices and the costs
for monitoring and other regulatory requirements. The costs of technology develop-
ments, the regulatory environment, the base for storage potential and etc. must be
added to the above expenditures. The main costs of a CCS project are divided into
three components including, capture and compression, transport and storage. Capture
and compression for large resources of CO2 like plants is the largest component of
the whole CCS costs. The most economical and common way of CO2 transportation
is through pipelines. Construction costs (material, labor, and etc.), operation and
maintenance costs (maintenance, monitoring, energy costs and etc.) and other costs
(insurance, design, fees and etc.) are the three main cost components for transporting
pipelines. Geological storage, ocean storage and storage via mineral carbonation are
three main methods for storing CO2 . In this section, the costs of storing in geological
formation have been considered. The cost estimates of storage in geological forma-
tion can be made with certainty because the equipment and technologies which use
in geological formation storage widely used in oil and gas industries. However, some
specific factors like onshore vs offshore, reservoir depth and geological characteris-
tics like thickness, permeability, and porosity create a considerable variability and
range of costs [66].
1 Introduction to Gas Injection 17

1.7.3 EOR

The costs of an EOR gas injection process is divided into investment and operation
costs. Investment costs include initial gas acquiring costs, compressors, injection
plants, gas injection facilities, production facilities and wells work-over, transmis-
sion pipelines, new production and injection wells and surface facilities. The most
expensive part of an EOR gas injection is building gas recovery plants.
Well treatments (corrosion inhibitors, scale inhibitors and paraffin inhibitors),
well workovers and servicing, power, labor, and facility maintenance are the most
important operating costs. A large percentage of field operating costs usually spent
on well workovers [114].

1.7.4 Underground Gas Storage

The costs of underground gas storage contain gas purchasing costs, well costs,
and dehydration and compression facilities. In the comparison of other methods
of storage, aquifers and depleted fields supply more storage capacity per invested
dollar. Storage in dry gas fields costs lower than in aquifers because of the lower
investment in wells and surface facilities [115].

References

1. Dudley B. BP statistical review of world energy 2017. London; 2017.


2. The role of natural gas in a sustainable energy market; 2017.
3. Zou C, Yang Z, Dai J, Dong D, Zhang B, Wang Y, Deng S, Huang J, Liu K, Yang C, Wei
G, Pan S. The characteristics and significance of conventional and unconventional Siniane
Silurian gas systems in the Sichuan Basin, central China. Mar Pet Geol. 2015:386–402.
4. Holditch SA. Unconventional oil and gas resource development–let’s do it right. J Unconv
Oil Gas Resour. 2013;1:2–8.
5. Agency IE. Resources to reserves 2013: oil, gas and coal technologies for the energy markets
of the future. International Energy Agency; 2013.
6. Dong Z, Holditch SA, McVay DA, Ayers WB. Global unconventional gas resource assessment.
Soc Pet Eng—Can Unconv Resour Conf 2011, CURC 2011 2011;1:810–25. https://doi.org/
10.2118/148365-ms.
7. Shukla R, Ranjith P, Haque A, Choi X. A review of studies on CO2 sequestration and caprock
integrity. Fuel. 2010;89:2651–64.
8. Fletcher PE. A review of gas-injection projects in West Texast. In: Drilling and production
practice. American Petroleum Institute; 1953.
9. Fowler FS. Emery West Pool gas-injection project. In: Drilling and production practice.
American Petroleum Institute; 1949.
10. Pings CJ Jr, Tempelaar-Lietz W. Canal field gas injection project. J Pet Technol. 1955;7:25–31.
11. Ramsey LA. Field histories show benefits of gas and water-injection projects in Eastern
Venezuela. J Pet Technol. 1960;12:33–6.
12. Leung DYC, Caramanna G, Maroto-Valer MM. An overview of current status of carbon
dioxide capture and storage technologies. Renew Sustain Energy Rev. 2014;39:426–43.
18 R. Azin et al.

13. Muggeridge A, Cockin A, Webb K, Frampton H, Collins I, Moulds T, et al. Recovery rates,
enhanced oil recovery and technological limits. Philos Trans R Soc A Math Phys Eng Sci.
2014;372:20120320.
14. Remme U, Blesl M, Fahl U. Global resources and energy trade: an overview for coal, natural
gas, oil and uranium; 2007.
15. Hayward T. BP statistical review of world energy 2010. London; 2010.
16. Dudley B. BP statistical review of world energy 2018. London; 2018.
17. Kuuskraa V, Stevens SH, Moodhe KD. Shale oil and shale gas resources: an assessment of
137 shale formations in 41 countries outside the US; 2013.
18. May R, Bata T, Schamel S, Fustic M, Ibatulin R. AAPG energy minerals division Bitumen
and heavy oil committee annual commodity; 2019.
19. McGlade C, Speirs J, Sorrell S. Unconventional gas—a review of regional and global resource
estimates. Energy. 2013;55:571–84. https://doi.org/10.1016/j.energy.2013.01.048.
20. Donaldson EC, Chilingarian GV, Yen TF. Enhanced oil recovery, II: processes and operations.
Elsevier; 1989.
21. BENNET E 0. Pressure maintenance. In: Drilling and production practice. Amarillo, Texas:
American Petroleum Institute; 1938.
22. Grow GC Jr. Survey of underground storage facilities in the United States. In: Drilling and
production practice. American Petroleum Institute; 1970.
23. Miller MG, Lents MR. Performance of Bodcaw reservoir, cotton valley field cycling project;
new methods of predicting gas-condensate reservoir performance under cycling operations
compared to field data. In: Drilling and production practice. American Petroleum Institute;
1946.
24. Justice WH. Review of cycling operations in the La Gloria field. J Pet Technol. 1950;2:281–6.
25. Green DW, Willhite GP. Enhanced oil recovery, vol. 6. Henry L. Doherty Memorial Fund of
AIME, Society of Petroleum Engineers Richardson, TX; 1998.
26. Koch HA Jr. High pressure gas injection is a success. World Oil. 1956;143:260–4.
27. Hall HN, Geffen TM. A laboratory study of solvent flooding; 1957.
28. Whorton LP, Brownscombe ER, Dyes AB. Method for producing oil by means of carbon
dioxide; 1952.
29. Whorton LP, Brownscombe ER. Method of recovering desirable petroleum hydrocarbon
fractions from producing oil reservoirs; 1955.
30. Whorton LP, Kieschnick WF. A preliminary report on oil recovery by high-pressure gas
injection. In: Drilling and production practice. American Petroleum Institute; 1950.
31. Stone HL, Crump JS. The effect of gas composition upon oil recovery by gas drive; 1956.
32. Barney CL. Petroleum vaporization recovery by high pressure gas injection. In: Fall meeting
of the Society of Petroleum Engineers of AIME. AIME, Society of Petroleum Engineers;
1957.
33. Welker JR. Physical properties of carbonated oils. J Pet Technol. 1963;15:873–6.
34. Patton JT, Coats KH, Spence K. Carbon dioxide well stimulation: part 1-A parametric study.
J Pet Technol. 1982;34:1, 791–798, 804.
35. Shah RP. A study of CO2 recovery and tertiary oil production enhancement in the Los Angeles
basin. US Department of Energy; 1979.
36. Horn FL, Steinberg M. Control of carbon dioxide emissions from a power plant (and use in
enhanced oil recovery). Fuel. 1982;61:415–22.
37. Hendriks CA, Blok K, Turkenburg WC. The recovery of carbon dioxide from power plants. In:
Climate and energy: the feasibility of controlling CO2 emissions. Springer; 1989. p. 125–42.
38. Marchetti C. On geoengineering and the CO2 problem. Clim Change. 1977;1:59–68. https://
doi.org/10.1007/bf00162777.
39. Cole CV, Flach K, Lee J, Sauerbeck D, Stewart B. Agricultural sources and sinks of carbon.
In: Wisniewski J, Sampson RN, editors. Terrestrial biospheric carbon fluxes quantification of
sinks and sources of CO2 . Dordrecht: Springer Netherlands; 1993. p. 111–22. https://doi.org/
10.1007/978-94-011-1982-5_7.
1 Introduction to Gas Injection 19

40. Blunt M, Fayers FJ, Orr FM Jr. Carbon dioxide in enhanced oil recovery. Energy Convers
Manag. 1993;34:1197–204.
41. Koide H, Tazaki Y, Noguchi Y, Nakayama S, Iijima M, Ito K, et al. Subterranean contain-
ment and long-term storage of carbon dioxide in unused aquifers and in depleted natural gas
reservoirs. Energy Convers Manag. 1992;33:619–26.
42. Gunter WD, Gentzis T, Rottenfusser BA, Richardson RJH. Deep coalbed methane in Alberta,
Canada: a fuel resource with the potential of zero greenhouse gas emissions. Energy Convers
Manag. 1997;38:S217–22.
43. Baklid A, Korbol R, Owren G. Sleipner Vest CO2 disposal, CO2 injection into a shallow under-
ground aquifer. In: SPE annual technical conference and exhibition. Society of Petroleum
Engineers; 1996.
44. Lyons WC, Plisga GJ. Standard handbook of petroleum and natural gas engineering. Elsevier;
2011.
45. Sandrea I, Sandrea R. Recovery factors leave vast target for EOR technologies. Oil Gas J.
2007;105:44–8.
46. Alagorni AH, Yaacob Z Bin, Nour AH. An overview of oil production stages: enhanced oil
recovery techniques and nitrogen injection. Int J Environ Sci Dev 2015;6:693–701.
47. Alfarge D, Wei M, Bai B. Comparative study for CO2 -eor and natural gases injection-
techniques for improving oil recovery in unconventional oil reservoirs. In: Carbon manage-
ment technology conference. Carbon Management Technology Conference; 2017.
48. Dong C, Hoffman BT. Modeling gas injection into shale oil reservoirs in the sanish field,
North Dakota. In: Unconventional resources technology conference. Society of Exploration
Geophysicists, American Association of Petroleum …; 2013. p. 1824–33.
49. Wang D, Zhang J, Butler R, Olatunji K. Scaling laboratory data surfactant imbibition rates
to the field in fractured shale formations. In: Unconventional resources technology confer-
ence. San Antonio, Texas, 20–22 July 2015, Society of Exploration Geophysicists, American
Association of Petroleum …; 2015. p. 2575–90.
50. Benalycherif M, Bishlawi M, Bencheikh A. Pressure maintenance in the Zarzaitine Field-
Algeria. 1972:21.
51. Wei MH, Yu JP, Moore DM, Ezekwe N, Querin ME, Williams LL. Case history of pressure
maintenance by Crestal gas injection in the 26R gravity drainage reservoir. In: SPE Western
regional meeting. Society of Petroleum Engineers; 1992. https://doi.org/10.2118/24035-MS.
52. van Wingen N, Barton CW, Case CH. Coalinga nose pressure maintenance project. J Pet
Technol. 1973;25:1147–52. https://doi.org/10.2118/4183-PA.
53. Torabzadeh J, Langnes GL, Robertson JO Jr, Yen TF, Donaldson EC, Chilingarian GV. Gas
injection. Enhanc Oil Recover II Process Oper. 1989:91.
54. Dong M, Huang S. Flue gas injection for heavy oil recovery. J Can Pet Technol 2002;41.
55. Al-Hussainy R, Ramey HJ Jr, Crawford PB. The flow of real gases through porous media. J
Pet Technol. 1966;18:624–36.
56. Alvarado V, Manrique E. Enhanced oil recovery: an update review. Energies. 2010;3:1529–75.
57. Moreno JE, Gurpinar OM, Liu Y, Al-Kinani A, Cakir N. EOR advisor system: a comprehensive
approach to EOR selection. In: International petroleum technology conference. International
Petroleum Technology Conference; 2014.
58. Azin R. Investigation of the VAPEX process in fractured reservoirs. Ph.D. dissertation Sharif
University of Technology; 2007.
59. Azin R, Kharrat R, Ghotbi C, Rostami B, Vossoughi S. Simulation study of the VAPEX
process in fractured heavy oil system at reservoir conditions. J Pet Sci Eng. 2008;60:51–66.
60. Azin R. The vapor extraction (VAPEX) process in heavy oil fractured systems. Ger Lambert
Academic Publishing; 2012.
61. AzIN R, Kharrat R, Ghotbi C, Vossoughi S. Improved heavy oil recovery by VAPEX process
in the presence of vertical and horizontal fractures. J Japan Pet Inst. 2007;50:340–8.
62. Azin R, Kharrat R, Vossoughi S, Ghotbi C. Study of the VAPEX process in fractured physical
systems using different solvent mixtures. Oil Gas Sci Technol l’IFP. 2008;63:219–27.
20 R. Azin et al.

63. Azin R, Kharrat R, Ghotbi C, Vossoughi S. Applicability of the VAPEX process to Iranian
heavy oil reservoirs. In: SPE Middle East oil gas show conference. Society of Petroleum
Engineers; 2005.
64. Izadpanahi A, Azin R, Osfouri S, Malakooti R. Optimization of two simultaneous water and
gas injection scenarios in a high GOR Iranian oil field. In: 82nd EAGE Annual conference &
exhibition, vol. 2020. European Association of Geoscientists & Engineers; 2020. p. 1–5.
65. Izadpanahi A, Azin R, Osfouri S, Malakooti R, GhanaatpishehSenani E. Production optimiza-
tion in a fractured carbonate reservoir with high producing GOR. Arab J Geosci. 2021; Under
Revi.
66. Metz B. Carbon dioxide capture and storage: IPCC special report. Summary for policymakers:
a report of working group III of the IPCC. Technical summary: a report accepted by working
group III of the IPCC but not approved in detail. Intergovernmental panel on climate change;
2005.
67. Azin R, Malakooti R, Helalizadeh A, Zirrahi M. Investigation of underground sour gas storage
in a depleted gas reservoir. Oil Gas Sci Technol d’IFP Energies Nouv. 2014;69:1227–36.
68. Malakooti R, Azin R. Simulation study of underground gas storage. Ger Lambert Academic
Publishing; 2012.
69. Malakooti R, Azin R. The optimization of underground gas storage in a partially depleted gas
reservoir. Pet Sci Technol. 2011;29:824–36.
70. Malakooti R, Azin R. Simulation study of underground gas storage. Lambert Academic
Publishing; 2012.
71. Adel H, Tiab D, Zhu T. Effect of gas recycling on the enhancement of condensate recovery,
case study: Hassi R’Mel South field, Algeria. In: International oil conference and exhibition
in Mexico. Society of Petroleum Engineers; 2006.
72. Taber JJ, Martin FD, Seright RS. EOR screening criteria revisited-Part 1: introduction to
screening criteria and enhanced recovery field projects. SPE Reserv Eng. 1997;12:189–98.
73. Taber JJ, Martin FD, Seright RS. EOR screening criteria revisited—part 2: applications and
impact of oil prices. SPE Reserv Eng. 1997;12:199–206.
74. Simpson MR. The CO2 huff’n’puff process in a bottomwater-drive reservoir. J Pet Technol.
1988;40:887–93.
75. Leung LC. Numerical evaluation of the effect of simultaneous steam and carbon dioxide
injection on the recovery of heavy oil. J Pet Technol. 1983;35:1, 591, 599.
76. Haines HK, Monger TG. A laboratory study of natural gas HuffN’Puff. In: CIM/SPE
international technical meeting. Society of Petroleum Engineers; 1990.
77. Haskin HK, Alston RB. An evaluation of CO2 huff’n’puff tests in Texas. J Pet Technol.
1989;41:177–84.
78. Schenewerk PA, Thomas J, Bassiouni ZA, Wolcott J. Evaluation of a South Louisiana CO2
Huff’n’Puff field test. SPE/DOE enhanced oil recovery symposium. Society of Petroleum
Engineers; 1992.
79. Monger TG, Coma JM. A laboratory and field evaluation of the CO2 huff’n’puff process for
light-oil recovery. SPE Reserv Eng. 1988;3:1, 161–168, 176.
80. Cotter WH. Twenty-three years of gas injection into a highly undersaturated crude reservoir.
J Pet Technol. 1962;14:361–5.
81. Stewart FM, Garthwaite DL, Karebill FK. Pressure maintenance by inert gas injection in the
high relief Elk basin field. 1955;204.
82. Kelly P, Kennedy SL. Thirty years of effective pressure maintenance by gas injection in the
Hilbig field. J Pet Technol. 1965;17:279–81. https://doi.org/10.2118/946-PA.
83. Waller RL. The Northwest Avard pressure maintenance project. J Pet Technol. 1971;23:1074–
80. https://doi.org/10.2118/3156-PA.
84. Babson EC. A review of gas injection projects in California. In: SPE California regional
meeting. Society of Petroleum Engineers; 1989.
85. Vowel WR. Gas injection in Eastern Venezuela field increases oil recovery. J Pet Technol.
1958;10:31–3.
1 Introduction to Gas Injection 21

86. Abel W, Jackson RF, Wattenbarger RA. Simulation of a partial pressure maintenance gas
cycling project with a compositional model, Carson Creek field, Alberta. J Pet Technol.
1970;22:38–46. https://doi.org/10.2118/2580-PA.
87. Moradi B, Tangsiri Fard J, Rasaei MR, Momeni A, Bagheri MB. Effect of gas recycling on
the enhancement of condensate recovery in an Iranian fractured gas/condensate reservoir. In:
Trinidad and Tobago energy resources conference. Society of Petroleum Engineers; 2010.
https://doi.org/10.2118/132840-MS.
88. Jing T, Xiao L, Zhao J. Field study of enhancing oil recovery by gas recycling injection in
ultra deep heavy oil reservoirs. In: SPE Asia Pacific oil gas conference exhibition. Society of
Petroleum Engineers; 2008.
89. Kumar J, Agrawal P, Draoui E. A case study on miscible and immiscible gas injection pilots
in Middle East carbonate reservoir in an offshore environment. In: International petroleum
technology conference; 2015.
90. Gunawan S, Caie D. Handil Field: three years of lean gas injection into waterflooded reservoirs.
In: SPE Asia Pacific improved oil recovery conference. Society of Petroleum Engineers; 1999.
91. Owens H Jr, Kirk E Jr, Brinkley TW. Handling high pressure cycling projects at Northwest
Branch, Louisiana. J Pet Technol. 1959;11:31–4.
92. Tang X, Wang R, Zhang H. Innovative field-scale application of injecting condensate gas and
recycling gas into medium oil pool: a case study in Sudan. In: SPE enhanced oil recovery
conference. Society of Petroleum Engineers; 2013.
93. Alford ME. A review of the Arun field gas production/cycling and LNG export project. In:
Middle East oil technical conference and exhibition. Society of Petroleum Engineers; 1983.
94. Wichert E. The windfall field cycling project. J Can Pet Technol. 1963;2:6–8.
95. Thompson FR, Thachuk AR. Compositional simulation of a gas-cycling project, Bonnie Glen
D-3A Pool, Alberta, Canada. J Pet Technol. 1974;26:1, 281–285, 294.
96. Belaifa E, Tiab D, Dehane A, Jokhio S. Effect of gas recycling on the enhancement of
condensate recovery in Toual Field Algeria, a case study. In: SPE production and operations
symposium. Society of Petroleum Engineers; 2003.
97. Cobanoglu M, Khayrutdinov F, Linthorst S, Iqbal M. Improving condensate recovery of a
rich sour gas condensate field by gas recycling. In: SPE EOR conference oil gas West Asia,
Society of Petroleum Engineers; 2014.
98. Talukdar S, Rasmussen J, Wennberg KE, Lien LS. Reservoir management challenges and
improved hydrocarbon recovery activities on Kvitebjørn HPHT gas condensate reservoir. In:
EUROPEC/EAGE conference and exhibition. Society of Petroleum Engineers; 2009.
99. Manrique EJ, Thomas CP, Ravikiran R, Izadi Kamouei M, Lantz M, Romero JL, et al.
EOR: current status and opportunities. In: SPE improved oil recovery symposium. Society of
Petroleum Engineers; 2010.
100. Vogel JL, Yarborough L. The effect of nitrogen on the phase behavior and physical properties
of reservoir fluids. In: SPE/DOE enhanced oil recovery symposium. Society of Petroleum
Engineers; 1980.
101. Siregar S, Hagoort J, Ronde H. Nitrogen injection vs. gas cycling in rich retrograde condensate-
gas reservoirs. In: International meeting on petroleum engineering. Society of Petroleum
Engineers; 1992.
102. Moses PL, Wilson K. Phase equilibrium considerations in using nitrogen for improved
recovery from retrograde condensate reservoirs. J Pet Technol. 1981;33:256–62.
103. Peterson AV. Optimal recovery experiments with N2 and CO2 . Pet Eng Int 1978;50.
104. Fraim ML, Moffitt PD, Yannimaras DV. Laboratory testing and simulation results for high
pressure air injection in a waterflooded North Sea oil reservoir. In: SPE annual technical
conference and exhibition. Society of Petroleum Engineers; 1997.
105. Klara SM, Srivastava RD, McIlvried HG. Integrated collaborative technology development
program for CO2 sequestration in geologic formations—United States Department of Energy
R&D. Energy Convers Manag. 2003;44:2699–712.
106. Ahmadi MA, zeinali Hasanvand M, Shokrolahzadeh S. Technical and economic feasibility
study of flue gas injection in an Iranian oil field. Petroleum 2015;1:217–22.
22 R. Azin et al.

107. Shokoya OS, Mehta SAR, Moore RG, Maini BB, Pooladi-Darvish M, Chakma A. The
mechanism of flue gas injection for enhanced light oil recovery. J Energy Resour Technol.
2004;126:119–24.
108. Liu R, Zhang J, Meng L, Liu F, Zuo C. Feasibility study of steam/flue gas mixture injection
in low permeability reservoir. In: SPE project and facilities challenges conference at METS.
Society of Petroleum Engineers; 2011.
109. Suhendro S. Review of 20 years hydrocarbon gas cycling in the Arun Gas field. In: SPE/IATMI
Asia Pacific oil gas conference exhibition. Society of Petroleum Engineers; 2017.
110. Harvey H, Henry RL. A laboratory investigation of oil recovery by displacement with carbon
dioxide and hydrogen sulfide; 1977.
111. Lv S, Liao X, Chen H, Chen Z, Lv X, Zhou X. Predicting the effects of acid gas on enhanced
oil recovery in hydrocarbon gas injection. In: SPE Western regional meeting. Society of
Petroleum Engineers; 2016.
112. Kenefake D, Hawrysz D, Al-Mohannadi A. Large-scale acid gas injection facilities for gas
disposal. In: International petroleum technology conference; 2007.
113. Patton EC Jr. Evaluation of pressure maintenance by internal gas injection in volumetrically
controlled reservoirs. Trans AIME. 1947;170:112–55.
114. Sheng J. Enhanced oil recovery field case studies. Gulf Professional Publishing; 2013.
115. Coats KH. Some technical and economic aspects of underground gas storage. J Pet Technol
1966;18:1, 561, 566.
Chapter 2
PVT of Gas Injection

Reza Azin, Amin Izadpanahi, and Shahriar Osfouri

Abstract This chapter introduces the PVT challenges of gas injection. First, the
phase diagram of various gases, oil samples and mixtures of oil and gas are inves-
tigated. In the next part, the important PVT experiments are discussed in details.
These experiments consist of CCE (constant composition expansion), DL (differ-
ential liberation), CVD (constant volume depletion), flash separation and swelling
test. Some calculation must be considered for designing the gas injection process as
it is discussed in Sect. 2.4. In Sects. 2.5, 2.6 and 2.7, three cases are studied about
change of phase behavior due to gas injection, MMP calculation of gas injection
and asphaltene precipitation due to gas injection, respectively. The optimum design
of gas injection and PVT challenges associated with gas injection are presented in
Sects. 2.8 and 2.9, respectively.

Nomenclature

Acronyms

GIIP Gas Initially in place


MMP Minimum Miscibility Pressure

R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
S. Osfouri
Department of Chemical Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: Osfouri@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 23


R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_2
24 R. Azin et al.

Variables

fs Fugacity
K Dispersion coefficient
L Reservoir length
m Mass of initial gas in-place, lb
Mw Molecular Weight, g/gmol
p Pressure, psi
R Ideal Gas Constant, ft3 .psi/lbmol.R
S Saturation
SEO Enriched oil slug size
Ss Solvent slug size
t Time
T Temperature, °F
V Volume, m3 or ft3
vs Molar volume of solid phase
x Mole fraction
z Compressibility factor

Subscripts

bt Breakthrough
d Dew-point
EO/O Enriched-oil and oil
EO/S Enriched-oil and solvent
i Initial
inj Injected
l Liquid
O/S Oil-Solvent
origin Original
rel Relative
sat Saturation
S/CG Solvent and chase gas

2.1 Introduction

Understanding the thermodynamic interaction between the injection gas and the orig-
inal reservoir fluid is a fundamental step in gas injection. Distribution of components
in phases is dependent on equilibrium conditions reflected by equality of chemical
potential for each component in each phase. However, the main challenge in any
2 PVT of Gas Injection 25

gas injection operation is that thermodynamic equilibrium is rarely met due to high
rate of injection and limited contact time between injected gas and reservoir fluid.
Despite this fact, many gas injection processes are designed by assuming a thermo-
dynamic equilibrium is happening. Subsequently, the process is overdesigned to take
into account the uncertainties arisen by deviation from equilibrium conditions.
Co-injection of contaminants such as CH4 , H2 S, N2 , O2 and SO2 with CO2 stream
as acid gas and flue gas provides an extra benefit to operators and let them handle
greenhouse gas emission from reservoirs or plants in addition to miscible EOR and
CO2 sequestration projects. However, each contaminant may increase or decrease in
MMP. The combined effect of components depends on the composition of injection
gas and needs be examined and modeled.
This chapter introduces the PVT challenges of gas injection. First, the phase
diagram of various gases, oil samples and mixtures of oil and gas are investigated. In
the next part, the important PVT experiments are discussed in details. These exper-
iments consist of CCE, DL, CVD, flash separation and swelling test. Some calcu-
lation must be considered for designing the gas injection process as it is discussed
in Sect. 2.4. In Sect. 2.5, 2.6 and 2.7, three cases are studied about change of phase
behavior due to gas injection, MMP calculation of gas injection and asphaltene
precipitation due to gas injection, respectively. The optimum design of gas injection
and PVT challenges associated with gas injection are presented in Sects. 2.8 and 2.9,
respectively.

2.2 Phase Diagram

When a gas stream is injected into an oil reservoir, it mixes with the oil in the gas-oil
front. Also, the gas composition approaches an equilibrium state assuming infinite
mixing given enough time. This, however, does not mean that the bulk gas stream is
in equilibrium with the bulk oil in reservoir. In this case, thermodynamics is a useful
tool for phase behavior study and modeling injecting gas and reservoir oil system at
the mixing zone.
Different modes of injection in terms of miscibility, i.e. miscible, immiscible,
near-miscible, first-contact miscible (FCM), and multiple-contact miscible (MCM)
gas injection, are referred to phase behavior study and modeling of gas injection
for EOR purposes. The state of miscibility depends on the composition of injection
gas and reservoir oil, as well as operating temperature and pressure. In terms of the
phase envelope, the PT diagrams of injection gas and reservoir oil are getting closer
to each other as the degree of miscibility increases. The term “bubble point” stands
for a pressure at which the first bubble of vapor gas evolves from liquid hydrocarbon
at reservoir temperature.
This point is important in reservoir engineering, as it is the intermediate between
single and two-phase regions. Practically, the actual bubble point of a reservoir hydro-
carbon mixture may occur at a pressure above bubble point measured experimentally,
as the gas evolves from oil forms as micro volumes or smaller size; the volumes of
26 R. Azin et al.

gas need to grow enough to shape a visible “bubble”. Similar to bubble point, the term
“dew point” applies mostly to light hydrocarbon mixtures, gas and gas condensate
reservoirs. This term refers to a pressure at which the first drop of liquid hydro-
carbon appears at reservoir temperature. Also, the formation of liquid hydrocarbon
initiates at a pressure above the experimentally measured dew point, long before the
liquid droplet or foggy atmosphere is visible by the operator from sight glass of a
PVT cell. Both bubble point and dew point pressures are determined from thermody-
namic phase equilibrium calculations. The phase diagram is often used to determine
the type of reservoir fluid. Based on the oil composition, its phase diagram, and
reservoir temperature and pressure conditions, the hydrocarbon reservoir may be
distinguished as black oil, near-critical oil, volatile oil, gas condensate, wet gas or
dry gas.

2.2.1 Hydrocarbon Gas

Typical phase envelopes for different hydrocarbon injection gases are shown in
Fig. 2.1. Details of injection gas compositions for samples of this figure are given
in Table 2.1. Composition of light hydrocarbon (CH4 ) is dominant in all samples. A
typical phase diagram is comprised of two parts, bubble point line (curve) and dew
point line (curve).

1400

Sample 1
1200 Sample 2
Sample 3
1000 Sample 4
Sample 5
Pressure (Psi)

800

600

400

200

0
-320 -270 -220 -170 -120 -70 -20 30
Temperature (F)

Fig. 2.1 Typical PT diagram for different injection gases with composition given in Table 2.1
2 PVT of Gas Injection 27

Table 2.1 Composition of typical hydrocarbon-dominant injection gases


Component Sample 1 Sample 2 Sample 3 Sample 4 Sample 5
C1 90 85 80 85 80
CO2 5 10 15 5 5
H2 S 5 5 5 10 15

2.2.2 Non-hydrocarbon Gas

Table 2.2 shows the composition of non-hydrocarbon gas which the relevant phase
diagram is shown in Fig. 2.2. Compared to hydrocarbons, the phase diagram of CO2 -
rich injection gas differs significantly from hydrocarbon-based gas. The PT diagram
of CO2 -rich gas shifts towards right compared to CH4 -rich samples.

Table 2.2 Composition of typical CO2 -dominant injection gases


Component Sample 6 Sample 7 Sample 8 Sample 9 Sample 10
C1 20 10 10 5 5
CO2 80 80 90 90 95
H2 S 0 10 0 5 0

1400
Sample 6

1200 Sample 7
Sample 8

1000 Sample 9
Pressure (psi)

Sample 10

800

600

400

200

0
-200 -150 -100 -50 0 50 100
Temperature (F)

Fig. 2.2 Typical PT diagram for different injection gases with composition given in Table 2.2
28 R. Azin et al.

2.2.3 Oil Samples

Table 2.3 presents the composition of typical reservoir oil which are from the Iranian
oil reservoirs. Petroleum is a complex mixture of hydrocarbon compounds, some
non-hydrocarbon compounds like CO2 , H2 S, N2 , H2 O, noble gases, and some hybrid
compounds called oganometallic groups. Understanding an exact composition of a
reservoir oil helps to predict its properties. However, the exact analysis of crude oil
is practically impossible, as it contains hundreds to thousands of compounds. The
major hydrocarbon families of petroleum include paraffin (alkane), naphthene (cyclo-
alkane), aromatic, resin, and asphaltene. Its chemical composition varies from one
reservoir to another, depending on the source and origin of kerogen, the basic hydro-
carbon complex which further converts into oil, migration, depth and geothermal
conditions. Based on the concentration of acid gas components (CO2 , H2 S), the oil
may be considered sweet (having little or no acid gas or sulfur-containing component)
or sour (having large amounts of acid gas or sulfur-containing components).
Figure 2.3 presents the PT diagram for different oil with composition given in
Table 2.3.

Table 2.3 Composition of typical reservoir oil


Sample 1 Sample 2 Sample 3
Component Mole% Component Mole% Component Mole%
H2 S 0.44 H2 S 0.63 CO2 1.22
N2 1.66 N2 1.85 CH4 38.63
CO2 2.19 CO2 1.87 C2 H6 6.97
CH4 48.43 CH4 49.07 C3 H8 4.77
C2 H6 5.38 C2 H6 5.82 IC4 1.8
C3 H8 3.91 C3 H8 4.1 NC4 3.98
IC4 1.03 IC4 1.21 IC5 1.5
NC4 1.87 NC4 2.68 NC5 1.62
IC5 0.85 IC5 0.74 C6+ 39.51
NC5 0.98 NC5 0.65
FC6 3.06 FC6 4.07
C7+ 30.2 FC7 3.62
FC8 2.05
FC9 2.78
FC10 1.85
FC11 1.53
C12+ 15.48
2 PVT of Gas Injection 29

4500
Sample 1
4000 Sample 2
Sample 3

3500

3000
Pressure (psi)

2500

2000

1500

1000

500

0
0 100 200 300 400 500 600 700 800 900
Temperature (F)

Fig. 2.3 Typical PT diagram for different oil samples with the composition given in Table 2.3

2.2.4 Mixtures of Oil and Gas

When gas is injected into an oil reservoir, the combined PT diagram between injection
gas and reservoir oil changes dynamically, as the volume ratio of oil and gas in contact
is varying. Complete mixing of injection gas and reservoir fluid rarely occurs in field
operation; thus, one may expect a set of phase envelopes rather than a unique one.
In other words, the composition at gas/oil interface approaches equilibrium and a
spectrum of compositions can be detected from bulk gas to gas/oil interface and
into bulk oil. As a result, a spectrum of dew point pressure from bulk gas to gas/oil
interface and a spectrum of bubble point pressure from gas/oil interface to the bulk
of the oil is formed corresponding to compositional changes upon gas injection. The
in-situ enrichment of injected gas upon multiple contacts with reservoir oil enables
the operator to use relatively cheap dry gas for injection and let the gas-oil system
approach miscibility at reservoir conditions. Depending on the equilibrium between
the injection gas and reservoir fluid, three different zones are formed. Figure 2.4
presents the zones that are formed due to gas injection.
An important point on the phase diagram of reservoir oil is that it is for orig-
inal composition only. Pressure decline upon oil production causes removal of light
hydrocarbon from reservoir oil, leading to a dynamic shift in the phase diagram.
Based on the physical phenomena occurring in a gas injection, three fluid phases
may form as well as a solid phase. Fluid phases include bulk gas (either as a gas
phase or supercritical fluid), bulk oil, and gas-oil mixing zone. Also, formation and
precipitation of asphaltene as solid phase is reported in some gas injection projects
30 R. Azin et al.

Injection
Bulk Gas/Bulk Bulk
Gas Reservoir Reservoir
Fluid Fluid
Interaction

Fig. 2.4 Formation of zones due to gas injection

[1–8]. The solid phase may be precipitated asphaltene, wax, or combination of asphal-
tene, wax and salt, which separate from reservoir oil upon contact and mixing with
injection gas and change in thermodynamic conditions of original fluid.

2.3 Pressure-Volume-Temperature (PVT) Experiments

Based on the type and purpose of gas injection, different PVT tests may be required
for a complete analysis of fluid phase behavior. In gas injection for pressure mainte-
nance, the interaction between injection gas and reservoir oil involves gas diffusion
and dissolution (solubility test) into oil at the gas-oil contact, the former is known as
transport phenomenon rather than thermodynamic property. Like pressure mainte-
nance, gas dissolution and diffusion is the main measured PVT and transport proper-
ties in the case of CO2 sequestration. However, usually a more comprehensive PVT
study is ordered before going through gas injection for EOR and UGS purposes to
fully understand fluid phase behavior and characterize reservoir oil and its interaction
with injecting gas. These tests include, but not limited to, CVD, BPP (or DPP), Flash,
Swelling, DL, CCE, miscibility tests, IFT reduction. For under-saturated reservoirs,
oil formation volume factor and solution gas ratio are also measured. A detailed
description and procedure of important PVT tests are presented here.

2.3.1 Constant Composition Expansion

Constant composition expansion is accomplished on the crude oil or gas condensate


samples to obtain the P-V relations of these systems. No hydrocarbon is removed
from the cell during this test, and the composition of the mixture in the cell remains
constant. This test is performed to reach the following experimental data:
a. Saturation pressure (bubble-point or dew-point pressure).
2 PVT of Gas Injection 31

b. Isotherm compressibility coefficient of the single-phase fluid above the satura-


tion pressure.
c. Gas phase compressibility factors.
d. Total hydrocarbon volume as a function of pressure.
The experimental procedure of this test is shown in Fig. 2.5 and describes as
follow:
Step 1. A hydrocarbon fluid sample (oil or gas) is placed in a visual PVT cell at
reservoir temperature and pressure bigger than the initial reservoir pressure.
Step 2. The pressure is decreased at constant temperature by discharging mercury
from the cell.
Step 3. The hydrocarbon volume is recorded.
Step 4. Steps 2 and 3 are repeated until the pressure of the cell is reached to the
intended pressure.
The volume in saturation pressure is chosen as the reference volume. The relative
volume in this test expressed as the ratio of the volume of hydrocarbon system to the
volume at saturation pressure:

V
Vr el = (2.1)
VSat

Fig. 2.5 Schematic of CCE experiments steps


32 R. Azin et al.

2.3.2 Differential Liberation

In differential liberation, the liberated gas from an oil sample during the pressure
reduction is removed from the cell before establishing the equilibrium with the liquid
phase. Composition of total hydrocarbon system is varying in this process. This test
is a representative of the separation process in the reservoir. Also, it can simulate
the flowing behavior of hydrocarbon systems above the critical gas saturation. The
experimental data attained from this test include:
a. Solution gas-oil ratio as a function of pressure.
b. Oil shrinkage as a function of pressure.
c. Remaining oil density as a function of pressure.
d. Evolved gas properties such as, composition, compressibility factor and the
specific gravity.
The procedure of this test is summarized in the following steps and is shown in
Fig. 2.6:
Step 1. The reservoir fluid sample is placed in a PVT cell at the reservoir
temperature.
Step 2. The cell is pressurized to saturation pressure by injection of mercury and
the volume of liquid is recorded.
Step 3. The cell is depressurized by removing the mercury from the cell.
Step 4. The liberated gas is discharged from the cell and the pressure is maintained
constant by reinjecting mercury.
Step 5. The volume of removed gas and remaining oil are measured.
Step 6. Step 4 and 5 are repeated in different pressures until the atmospheric
pressure.
Step 7. The final volume of remaining oil is referred as the residual oil.

Fig. 2.6 Schematic of DL experiments steps


2 PVT of Gas Injection 33

2.3.3 Constant Volume Depletion

Constant volume depletion generally is conducted on gas condensate fluid samples


and it is necessary for a prediction of pressure depletion performance. Determining
reserves, evaluating field separation methods, planning for increasing liquid recovery
and designing future operations need reliable and accurate pressure depletion perfor-
mance. These predictions can be done with the help of experimental data collected
from the CVD test. These important laboratory measurements which are provided
by the CVD test include:
a. Dew-point pressure
b. Variation of composition in gas phase
c. Compressibility factor in reservoir conditions
d. Original in-place hydrocarbons recovery at any pressure
e. Accumulation of retrograde condensate
CVD test consists some steps which are shown in Fig. 2.7 and summarized as
follow:
Step 1. A specific amount of representative reservoir fluid is charged to the visual
PVT cell. The overall composition (zi ) of this fluid must be known. PVT cell is
maintained at reservoir conditions (reservoir pressure and temperature). Reference
volume is defined as the initial volume (Vi ) of the saturated fluid.
Step 2. The initial gas compressibility factor and Gas Initially In place are
calculated by:

Fig. 2.7 Schematic of CVD experiments steps


34 R. Azin et al.

pd Vi
zd = (2.2)
n i RT

G I I P = 379.4n i (2.3)

where
pd = dew-point pressure, psia.
Vi = initial gas volume, ft3 .
ni = initial number of moles of the gas = m/MWa .
m = mass of initial gas in-place, lb.
R = gas constant, 10.731.
zd = compressibility factor at dew-point pressure.
GIIP = Gas Initially In place.
Step 3. The cell pressure is decreased to a specified pressure (P). The pressure
variations are achieved by add or remove the mercury from the cell which illus-
trated in Fig. 2.7b. Decreasing the pressure causes the formation of a liquid phase.
When the equilibrium is established between the fluids in the cell, the total volume
(Vt ) and retrograde liquid volume (Vl ) are visually measured. Retrograde liquid
saturation is defined as the ratio of the volume of retrograde liquid to the initial
volume and mathematically shows as:

Vl
Sl = × 100 (2.4)
Vi

Step 4. The gas is removed while an amount of mercury reinjected to the cell
to maintain the cell pressure (at the constant pressure P) and to reach the initial
volume. This process is shown in Fig. 2.7c. The volume of the removed gas is
measured at the cell conditions.
Step 5. The properties and composition of the removed gas are determined. Also,
the volume of removed gas is measured in standard conditions.
These steps are repeated until the minimum test pressure is attained after which
the composition and quantity of the liquid and gas are measured.

2.3.4 Flash Separation

Flash separation is also known as the separator test which is conducted to evaluate the
volumetric behavior as the reservoir fluid passes through separators and then into the
stock tank. The results of this test are used to design the optimum surface facilities
and conditions. The combination of this test and differential liberation test data can
obtain the PVT parameters (Bo , Rs and Bt ) which are necessary for engineering
2 PVT of Gas Injection 35

calculations. The details of this test are not reported as this test is not involving in
the gas injection process.

2.3.5 Swelling Test

The swelling test is related to the reservoirs under gas injection or dry gas cycling
scheme. This test determines the dissolution of the injected gas in the reservoir
sample. It can be concluded from this test, each addition of injection gas causes the
relative volume and saturation pressure to increase. The following experimental data
are obtained from this test:
a. The relation between saturation pressure and the amount of injected gas.
b. The variation in fluid volume due to contact with the injected gas.
Swelling test experimental procedure is summarized in the following steps and
schematically is shown in Fig. 2.8:
Step 1. A visual PVT cell is filled with a reservoir fluid sample with a known
overall composition (zi ). The cell is maintained at reservoir temperature and
fluid saturation pressure. The volume of the fluid in this condition is recorded
as (Vsat )origin .
Step 2. A predesignate volume of proposed injection gas with known composition
is added to the cell. The cell pressure is increased until only one phase is present.
This pressure is defined as the new saturation pressure (ps ). Also, the new fluid
volume (Vsat ) is recorded in this condition. The relative volume in this test is

Fig. 2.8 Schematic of DL experiments steps


36 R. Azin et al.

defined as the ratio of the volume of fluid in new saturation pressure (Vsat ) to the
original fluid volume (Vsat )origin and mathematically shows:

Vsat
Vr el = (2.5)
(Vsat )orig

Step 3. Step 2 is repeated until the mole percent of injected gas attains a preset
value.

2.4 PVT Calculation in Gas Injection Processes

2.4.1 Equation of State

Results of PVT tests are used to tune equations of state (EOS), which are powerful
tools for mathematical modeling of phase behavior and thermodynamic studies for
reservoir fluid and their mixtures with injection gases. The common types of EOS in
reservoir engineering are cubic for compressibility factor or volume, which originate
from the well-known Van der Waals cubic EOS developed in 1873. Development
of this EOS was a turning point in modeling gas phase behavior and included the
role of attractive and repulsive forces in the ideal gas equation. Many modifications
were imposed on the parameters of Van der Waals-type EOS during the past century
to suggest accurate mathematical model of thermodynamic data and phase behavior
studies for a wide range of fluid systems. A general form of cubic EOS may be
written as follow.

f (P, V, T ) = 0 (2.6)

Parameters of this general form are then customized for any specific cubic EOS
[9]. Table 2.4 shows a series of EOS known as cubic EOS.
When tuned, the EOS is used to model and simulate phase envelope and PVT
experiments. As mentioned, measured PVT data are used to optimize the EOS
parameters in a systematic procedure known as EOS tuning. For a proper tuning

Table 2.4 Series of EOS


Reference Equation of state
Redlich_Kwong (1949) p= RT
V −b − a
V (V +b)T 0.5
a(T )
Soave_Redlich_Kwong (1972) p= RT
V −b − V (V +b)
a(T )
Peng_Robinson (1976) p= RT
V −b − V (V +b)+b(V −b)
a(θ,T )
Stryjek_Vera_Peng_Robinson p= RT
V −b − V (V +b)+b(V −b)
(1986)
a(θ,T )
Patel_Teja (1982) p= RT
V −b − V (V +b)+c(V −b)
2 PVT of Gas Injection 37

approach, it is assumed that PVT experiments are performed at controlled condi-


tions and fluid samples are consistent and their mixture represents original reservoir
fluid. However, such conditions may not exist in some cases. Any failure in proper
quality check (QC) of selected fluids or deviation from equilibrium conditions may
result in erroneous PVT modeling [10, 11]. The first steps in PVT studies is taking
fluid samples and recombining them to obtain original reservoir fluid. These steps
are shown schematically in Fig. 2.9.
Osfouri and Azin reviewed the challenges associated with sampling and recom-
bination of hydrocarbon fluids [10]. According to their study, main sources of
error during sampling include separator instability, lack of vapor-liquid equilibrium,
volume ratio of separator outlets, and the presence of contaminants. Sometimes,
analysis of separator outlet streams indicates the deviation from equilibrium for
some wells which can have direct impact on saturation pressure and oil (condensate)
to gas ratio prediction of well stream. Also, improper fluid handling, oil and gas
leakage, and the presence of corrosive compounds can severely affect the quality
of the recombined fluid. Overall, Thermodynamic equilibrium is a key factor for
a successful sampling. Osfouri et al. proposed an algorithm for QC of PVT data
[12]. Schematic of this algorithm is shown in Fig. 2.10. The first step of PVT QC is
checking the validity of the samples of oil and gas for recombination. The second
step is the verification thermodynamic equilibrium conditions of the samples. The
Rs should be corrected in the case that oil carries over in the gas stream in the third
step. In the next step, the material balance should be checked for the oil and gas
phases. Finally, for gas condensate samples the validity of CVD tests is checked. In
the case of negative composition for one or more components, the compositions that
reported in CVD test should not be used tuning the EOS and regression [12].

Fig. 2.9 Schematic of fluid sampling and recombination [12]


38 R. Azin et al.

Fig. 2.10 Schematic of the algorithm for QC of PVT data [12]

Also, there are components in the fluid composition whose composition are not
clear; these “pseudo components” are lumped form of other pure components and
their physical and critical properties need to define before going through EOS tuning.
This procedure is known as “heavy end” or “plus fraction” fluid characterization
Sometimes, the users decide to lump selected components to reduce the total number
of components and avoid large runtime in reservoir simulation. The EOS tuning is
a nonlinear, multivariable optimization problem which seeks for optimized EOS
parameters to fit best with measured PVT data and predict accurate thermodynamic
properties. The EOS tuning has been a topic of interest for chemical and petroleum
engineers since 80s and nearly all field development studies considered comprehen-
sive studies. Osfouri et al. proposed an integrated strategy for fluid characterization
and EOS tuning [11]. Schematic of this integrated approach is shown in Fig. 2.11
According to this approach, either dynamic (D) or static (S) strategy is described by a
“path” which specifies the class of fluid or SCN of plus fraction, splitting and lumping
(if any), type of empirical correlation, and grouping step (in dynamic strategy). For
class B, Splitting C12+ to C12 and C13+ was done by either exponential (EX) or gamma
(GA) distribution function. For example, the path named as S-B-GA-L refers to a
case in static strategy in which grouping class B with gamma distribution function
and Lee-Kesler [13] correlation is studied. The abbreviation “NO” in classes A and
2 PVT of Gas Injection 39

Fig. 2.11 Schematic of an integrated approach for fluid characterization and EOS tuning [11]

C indicate that no splitting was made on them and both are run with composition up
to C12+ . For dynamic strategy, for example, D7-EX-L-1 refers to a case in which C7+
is split by exponential distribution function and Lee-Kesler [13] correlation is used
to characterize SCN and lumped groups. The last number (1 in this example) refers
to the grouping step of Whitson’s procedure [6].
Therefore, proper EOS tuning is highly dependent on the quality of measured PVT
data and fluid characterization. However, there are debates on the overall process of
EOS tuning and fluid characterizations in gas injection. First, the PVT experiments
are conducted at controlled and near-equilibrium conditions, while reservoir condi-
tions are rarely approached equilibrium. Second, a loop between invalid PVT data
and improper EOS tuning will result in improper fluid phase determination with
inherent errors.

2.4.2 Gas-Liquid Miscibility

Component transfer between phases is occurred through mass transfer due to the
lack of equilibrium. According to Danesh [9], miscibility and thermodynamic equi-
librium at the gas-liquid interface is just an assumption and may not be observed
in real-world field application. The equilibrium between phases can be violated by
accelerated interphase mass transfer. This transport phenomenon continues until
the composition of gas and liquid come into equilibrium and chemical potential of
components become equal at the gas-liquid front. If the properties of two phases
40 R. Azin et al.

are similar in terms of composition and intermolecular forces and interfacial tension
approaches zero, the two phases become miscible. Miscibility is a thermodynamic
state which is affected by composition of injection gas and reservoir oil, tempera-
ture, and pressure. At a given temperature, miscibility increases with pressure for
gas-hydrocarbon systems.

2.4.3 Different States of Miscibility

2.4.3.1 First-Contact Miscibility (FCM)

Miscibility is a dynamic process and may occur at the first contact between phases,
known as first-contact miscibility (FCM). In operation of this process, a relatively
small primary slug is injected which is miscible with reservoir fluid. Then a larger
and less expensive slug is followed the primary slug. The size of each slug depends
on the economics of operation. Also, the primary and secondary slug should reach
the miscibility conditions for an efficient FCM process [14]. Also, it is necessary to
inject enough solvent to ensure that FCM is maintained to the end of the process. A
simple dispersion model can determine the amount of required solvent to meet this
criterion. The following equation can be used to determine the minimum slug size
in the FCM process [15]:

4.652   
(Ss )min = K O/S × tbt + K S/C G × tbt (2.7)
2L
where
(Ss )min = minimum solvent slug size.
L = reservoir length.
KO/S = dispersion coefficient between oil and solvent.
KS/CG = dispersion coefficient between solvent and chase gas.
and tbt = breaking through time.
On the other hand, methane flooding at reservoir conditions is an immiscible flood
because methane and crude oil are partially soluble. But, crude oil can dissolve hydro-
carbons with higher molecular weight such as propane and LPG at most reservoir
conditions. In this case, LPG and propane become liquid when pressure and temper-
ature are increased. If the methane is brought to contact with propane at specific
conditions (above the critical point) propane changes from liquid to gas and become
soluble in methane. In this case, propane and methane will mix in all proportions
and formed a single phase gas [16].
2 PVT of Gas Injection 41

2.4.3.2 Multiple-Contact Miscibility (MCM)

For most systems, the composition of injection gas and reservoir oil is not miscible
at the first contact and gradually approach miscible conditions through successive
contacts, which is known as multiple-contact miscibility (MCM). Therefore, the
MCM refers to gas-oil systems which are immiscible initially, but approach misci-
bility upon successive contacts which may be achieved by vaporization of light
and intermediate components of reservoir oil into the gas phase, or condensation
of intermediate components of injection gas into reservoir oil. In the case of lean
gas injection, mass transfer of components between phases may lead to enrichment
of the gas phase with intermediate heavy components vaporized from liquid hydro-
carbon. In this case, the miscibility process is known as vaporizing drive miscibility
or vaporizing gas drive (VGD). Alternatively, when rich gas is injected into oil reser-
voir, the concentration gradient may force transfer of intermediate components from
the gas phase into oil, and intermediate components condense into the liquid phase.
This mechanism is known as condensing drive miscibility or condensing gas drive
(CGD). When both condensing and vaporizing mechanisms are active in a gas-oil
contact, the process is known as condensing/vaporizing gas drive (CVGD).
The effects of dispersion and dispersive mixing as well as the amount of required
solvent for enriching the oil must be considered for calculation the slug size for
a condensing solvent MCM displacement. Also, slug size must enriched oil by
providing enough intermediate molecular weight components. Slug size calculation
can be summarized in the following steps [15]:
Step 1. Determine the necessary amount of enriched oil to dominate the dilution
effect using the following equation:

4.652   
(S E O )min = K E O/O × tbt + K E O/S × tbt (2.8)
2L
where
(SEO )min = minimum enriched oil slug size.
L = reservoir length.
KEO/O = dispersion coefficient between enriched-oil and oil.
KEO/S = dispersion coefficient between enriched-oil and solvent.
and tbt = breaking through time.
Step 2. Calculate the amount of required solvent to create the volume of enriched
oil which calculated in step 1 using the equation of state in a phase behavior
software modelling and verified in the laboratory.
Step 3. Determine the volume of the required solvent to dominate the dilution
process using the following equation

4.652   
(Ss )min = K S/E O × tbt + K S/C G × tbt (2.9)
2L
42 R. Azin et al.

where
(Ss )min = minimum solvent slug size.
L = reservoir length.
KS/EO = dispersion coefficient between solvent and enriched-oil.
KS/CG = dispersion coefficient between solvent and chase gas.
and tbt = breaking through time.
Step 4. Add the volumes determined in the step 2 and 3.

2.4.4 Minimum Miscibility Pressure (MMP)

The lowest pressure at which miscible conditions prevail is known as minimum misci-
bility pressure (MMP). For pressures above MMP, the system is certainly miscible.
When the injection pressure high enough that the injected gas can become miscible
with reservoir oil at the first contact. This pressure is defined as the first contact
miscibility pressure (Pmax ) [17]. Temperature, oil composition, gas composition and
initial gas-oil ratio are the four important parameters that affect Pmax and MMP.
These pressures are increased almost linearly with increasing the temperature that
concludes reservoir temperature has a strong effect on the crude oil-CO2 miscibility
development. The effect of oil composition (dead and live oil) is weak in the injection
of pure CO2 . The effect of gas composition was studied by adding CH4 to the pure
CO2 which causes to increase the Pmax and MMP. Also, the injection of produced
gas can reach miscibility with reservoir oil at the actual reservoir conditions. The
initial GOR has a strong effect on the Pmax and MMP which depend on the amount
of CH4 component of the live oil-impure CO2 system [18].

2.4.5 MMP Measurements

Experimental techniques used for measuring miscibility conditions between phases


include slim tube test and rising bubble apparatus (RBA) [9]. A drawback of the slim
tube experiment is that physical dispersion is likely to occur and cause misleading
results [19]. Visual PVT Cell is also used for detecting MMP and miscibility condi-
tions at successive pressures. Johnson and Pollin described the process of MMP
determination using PVT cell [20]. They pointed out that MMP is associated with
critical pressure of gas/oil mixtures when pure hydrocarbon is used as oil. They
injected CO2 into pure alkanes and figured out that measured MMP was close to
critical pressure of the mixture. However, such a conclusion may not be valid in
determining MMP for a complex reservoir fluid. Apart from the experiment, MMP
may be determined through compositional simulation, mixing cell models, analytical
methods and empirical correlations.
2 PVT of Gas Injection 43

2.4.6 MMP Correlations

Many correlations have been proposed to predict MMP of different gases in crude
oil, some of which will be reviewed here. The correlations are developed for non-
hydrocarbon and hydrocarbon gases. In general, developed correlations are a function
of reservoir temperature, the composition of reservoir oil and injection gas, proper-
ties of oil heavy end (plus fraction), and the ratio of light to intermediate components
in reservoir oil. Some correlations consider proportions and properties of interme-
diate components (C2 –C6 ) [21] and C5+ properties [21, 22] The mole% of volatile
components (C1 , N2 ) and volume ratio of volatile to intermediate components are
also considered in some correlations [22]. A critical review, evaluation and compar-
ison of MMP correlations is addressed by Shokir [23, 24], Emera and Sarma [25],
and Yuan et al. [19].

2.4.6.1 Non-hydrocarbon Gas Correlations

Alston et al. Correlation

Alston et al. developed MMP correlations for CO2 -live oil (LO) and CO2 -stock tank
oil (STO) [21].
CO2 -LO:

pC O2 −L O = 8.78 × 10−4 (TR )1.06 (MwC5+ )1.78 (xvol /xint )0.136 (2.10)

Which, TR : Reservoir Temperature, °F.


MC5+ : Molecular weight of C5+
xint : mole fraction of intermediate (C2 –C4 ), CO2 and H2 S in oil.
xvol : mole fraction of volatile (C1 , N2 ).
CO2 -STO:

pC O2 −ST O = 8.78 × 10−4 (TR )1.06 (MwC5+ )1.78 (2.11)

In this model, equation, N2 , CO2 are volatile components and C2 –C4 are interme-
diate components. The MMP is determined in psia. The authors extended their LO
correlation to predict MMP of impure streams of CO2 and other non-condensable
gases using a correction factor:

(M M PC O2 )imp−Lo = (M M PC O2 ) × Fimp (2.12)

F imp is the correction factor for impure CO2 :


 
  1.935×87.8
87.8 Tcm
Fimp = (2.13)
Tcm
44 R. Azin et al.

Table 2.5 Range of


Parameter Ranges
applicability for Alston et al.
[21] Temperature, °F [K] 90–243 [305–390]
Volatile/intermediate ratio 0.14–13.61
C5+ molecular weight 169.2–302.5
Weight average critical 70.7–115.7 [294.7–319.7]
temperature, °F [K]
Experimental MMP, psia 948–4930 [6.5–34]
[MPa]

In this equation, T cm is the pseudo critical temperature of gas mixture (°F) calcu-
lated by summing the component critical temperature, Tci (°R), multiplied by its
weight factor, wi :


n
Tcm = wi Tci − 459.7 (2.14)
i=1

Range of applicability for Alston et al. correlation [21] is given in Table 2.5.

Johnson and Pollin Correlation

Johnson and Pollin proposed an accurate and reliable empirical correlation for
predicting the minimum miscibility pressure. This correlation can be used for a
wide range of live oil and stock tank oil and pure or diluted CO2 as injection gas
[20].

PM M P − Pci = αi (T − Tci ) + I (β M − Mi )2 (2.15)

where PMMP and Pci are the miscibility pressure and the injection gas critical pres-
sure. T, Tc , M and Mi are the reservoir temperature, injection gas critical tempera-
ture, the average molecular weight of the oil and molecular weight of injection gas,
respectively. The other parameters of this equation are defined as follow:

I = C11 + C21 M + C31 M 2 + C41 M 3


+ (C12 + C22 M)A P I + C13 A P I 2 (2.16)

Constants of equations are listed in Table 2.6.


αi depends on the type of injecting gas:
 
psia
For pure CO2 : αi = 18.9 (2.17)
K
2 PVT of Gas Injection 45

Table 2.6 Constants of


C11 = − C21 = 0.06313 C31 = −1.954 C41 = 2.052
Eqs. 2.15 and 2.16
11.7300 × 10−7
C12 = C13 = −7.222 C22 = −1.138 β = 0.285
0.136 × 10−5 × 10−5

 
1000y2
For CO2 + N2 : αi = 10.51 1.8 + (2.18)
(T − Tci )
 
100y2
For CO2 + CH4 : αi = 10.5 1.8 + (2.19)
(T − Tci )

In this correlation, Pci , Tci , and Mi are critical pressure, critical temperature and
molecular weight of injection gas. M and API refer to molecular weight and API
gravity of reservoir oil.

Shokir Correlation

Shokir used a graphical user interface program called GRACE to develop his model.
The proposed MMP correlations depend on reservoir temperature, vol% of light and
intermediate hydrocarbons, the molecular weight of C5+ , and mole % of C1 , C2 –C4 ,
N2 , and H2 S in the injected gas [23, 24]. Details of Shokir correlation is as follows:

M M P = −0.068616 × Z 3 + 0.31733 × Z 2
+ 4.9804 × Z + 13.432 (2.20)


8
Z= Zn (2.21)
i=1

Z n = A3n xn3 + A2n xn2 + A1n x2 + A0 (2.22)

This correlation applies for pure and impure CO2 streams. The impurities include
CH4 , N2 , H2 S, and C2 –C6 . Constants of Shokir correlation are given in Table 2.7.

Kamari et al. Correlation

Kamari et al. correlation is developed based on Gene Expression Programming (GEP)


approach for CO2 -oil MMP. This correlation is a function of Tc , TR, vol/int (the ratio
of volatile to intermediate components)and MwC5+ in live oil system [26]. The details
of this correlation given as follow:
46 R. Azin et al.

Table 2.7 Constants of Shokir correlation [23]


n x A3 A2 A1 A0
1 TR 2.3660E−06 −5.5996E−04 7.5340E−02 −2.9182E+00
2 %vol −1.3721E−05 1.3644E−03 −7.9169E−03 −3.1227E−01
3 %interm 3.5551E−05 −2.7853E−03 4.2165E−02 −4.9485E−02
4 MwC5+ −3.1604E−06 1.9860E−03 −3.9750E−01 2.5430E+01
5 %C1 1.0753E−04 −2.4733E−03 7.0948E−02 −2.9651E−01
6 %C2 −%C4 6.9446E−06 −7.9188E−05 −4.4917E−02 7.8383E−02
7 %N2 0 3.7206E−03 1.9875E−01 −2.5014E−02
8 %H2 S 3.9068E−06 −2.7719E−04 −8.9009E−03 1.2344E−01

A4
M M P = 1.138A1 A2 −
1.1384 A1 A2 A3
2.6354
+
A1 A2 A3 (1.138A1 A2 − A4
1.138A1 A2 A3
+ 0.0613)
+ 0.0613 (2.23)

where A1 , A2 , A3 , A4 and A5 are defined as follow:

A1 = A5 + A25 − 0.3016 (2.24)

A2 = A25 − 0.3016 (2.25)

A4 2.8381
A3 =   (2MwC5+ +   )2
Tc2 V ol. V ol.
I nt. I nt.
− 2.6558 (2.26)

2MwC5+ 2.8381
A4 = +   
Tc T c V ol. I nt.
⎛ ⎞2
2MwC5+ 2.8381
+⎝ +   ⎠ (2.27)
Tc T c V ol. I nt.

TR MwC5+ A4
A5 = + −
TC TC TC A3
⎛ ⎞2
A4 2MwC5+ 2.8381
−   ⎝ +   ⎠
TC3 V ol. I nt. Tc T c V ol. I nt.
2 PVT of Gas Injection 47

2.6558
+ (2.28)
TC

Kaydani et al.

This correlation is developed using Multi-gene genetic programming (MGGP) [27].


Details of this correlation is as follow:
Kaydani et al. are reported that the correlation (2.29) is the best MGGP model
with the minimum errors for pure CO2 injection. This correlation is a function of
reservoir temperature, molecular weight of C5+ , and mole fraction of intermediate
and volatile components.

M M P = 0.2683 × TR + 0.2683 × MwC5+


+ 0.2683 × TR (TR − X int )2
+ 0.3339 × TR (TR + X vol − X int ) + 0.1161 (2.29)

Kaydani et al. also proposed a new correlation for Fimp as follows



4
X Cont × X C O2
Fimp = + 0.04 × X C O2 + 0.003 (2.30)
TR

where Xcont , XCO2 , and TR are the concentration contaminant components (N2 ,
C1 , H2 S, and C2 –C4 ), CO2 concentration and reservoir temperature. The value of
concentration must be entered based on 1 mol and normalized temperature.

Emera and Sarma Correlation

Emera and Sarma developed a correlation for estimating the MMP when CO2 is
mixed with other gases. This correlation is developed using genetic algorithm (GA)
[25], as follows:
 
pr,impureC O2 1.8TC W + 32
= 3.406 + 5.786 ×
pr, pur eC O2 1.8TC,C O2 + 32
   
1.8TC W + 32 2 1.8TC W + 32 3
− 23.0 × + 20.48 ×
1.8TC,C O2 + 32 1.8TC,C O2 + 32
 4
1.8TC W + 32
− 5.7 × (2.31)
1.8TC,C O2 + 32

where
M M PimpureC O2
pr,impureC O2 = (2.32)
pC W
48 R. Azin et al.

Table 2.8 Constants of Emera and Sarma correlation [25]


Components SO2 H2 S CO2 C2 C1 N2 Other components
MFi 0.3 0.59 1 1.1 1.6 1.9 1

Table 2.9 Constants of Yuan


MC7+ (g/mol) 139–319 a5 1.1667E−01
correlation for Pure CO2 [19]
PC2−6 (%) 2–40.3 a6 8.1661E+03
a1 −1.4634E+03 a7 −1.2258E−01
a2 0.6612E+01 a8 1.2283E−03
a3 −4.4979E+01 a9 −4.0152E−06
a4 0.2139E+01 a10 −9.2577E−04


n
pC W = wi pCi (2.33)
i=1

M M Ppur eC O2
pr, pur eC O2 = (2.34)
pC,C O2

n
TC W = M Fi wi Tci (2.35)
i=1

Values of MFi for various components are reported in Table 2.8.


In the case of MMP < oil bubble point pressure (pb ), pb is chosen as the MMP.

Yuan et al. Correlation

Yuan et al. reviewed and evaluated common CO2 -MMP correlations, and proposed
their correlation for pure and impure CO2 streams [19].
For pure CO2 :

M M Ppur e = a1 + a2 MwC7+ + a3 PC2−6


PC2−6
+ (a4 + a5 MC7+ + a6 )T
MwC7+
+ (a7 + a8 MwC7+ + a9 MwC
2
7+
+ a10 PC2−6 )T 2 (2.36)

where T, MwC7+ and PC2−6 are the reservoir temperature, C7+ molecular weight and
total molar percentage of C2 –C6 , respectively. The range of MC7+ , PC2−6 and the best
fit coefficient for Eq. (2.36) are reported in Table 2.9.
Impure CO2 :
2 PVT of Gas Injection 49

Table 2.10 Constants of


a1 −6.5996E−02 a6 −2.7344E−02
Yuan correlation for impure
CO2 [19] a2 −1.5246E−04 a7 −2.6953E−06
a3 1.3807E−03 a8 1.7279E−08
a4 6.2484E−04 a9 −3.1436E−11
a5 −6.7725E−07 a10 −1.9566E−04

M M Pimp
= 1 + m(PC O2 − 100) (2.37)
M M Ppur e

where

m = a1 + a2 MwC7+ + a3 PC2−6
PC2−6
+ (a4 + a5 MwC7+ + a6 )T
MwC7+
+ (a7 + a8 MwC7+ + a9 MwC
2
7+
+ a10 PC2−6 )T 2 (2.38)

The coefficients of Eq. (2.38) are reported in Table 2.10.


Equation (2.38) is strictly reliable only for methane contents in the gas up to 40%.

Glaso Correlation

Glaso’s correlation is suggested for the MMP calculation in miscible injection with
N2 . This correlation is a function of reservoir temperature, molecular weight of C7+
and the mole percent of methane and intermediate components in reservoir fluid [28].
Equation (2.39) is the proposed MMP correlation for the oil with the API lower than
40.

M M P = 80.14 + 35.35H + 0.76H 2 A P I < 40 (2.39)

0.88
MwC T 0.11
H= 7+
(2.40)
[(C2−6 )0.64 (C1 )0.33 ]

The proposed MMP correlation for the oil with the API higher than 40 is as
follows:

M M P = −648.5 + 2619.5H − 1347.6H 2 A P I > 40 (2.41)

0.48
MwC
H= 7+
(2.42)
[T 0.25 (C2−6 )0.12 (C1 )0.42 ]
50 R. Azin et al.

2.4.6.2 Hydrocarbon Gas Correlations

Glaso Correlation

Glaso presented a generalized MMP correlation which derived from graphical corre-
lations. This correlation can predict the MMP for MCM displacement by hydro-
carbon, N2 or CO2 gas and it is a function of reservoir temperature, molecular weight
of C7+ , mole percent of CH4 and the molecular weight of intermediates in the injection
gas [29]. This correlation is shown as:

M M P = 810 − 3.404MwC7+
−1.058
+ (1.7 × 10−9 MwC
3.730
7+
exp(786.8MwC 7+
))T
FR > 18% (2.43)

M M P = 2947.9 − 3.404MwC7+
−1.058
+ (1.7 × 10−9 MwC
3.730
7+
exp(786.8MwC 7+
))T
− 121.2FR FR < 18% (2.44)

FR is the mole percent of C2 –C6 in reservoir fluid. T is reservoir temperature in


°F, and MMP is expressed in psig.

Firoozabadi and Khalid Correlation

Firoozabadi and Khalid correlation can predict the MMP for VGD process. The
effective parameters in this correlation are the ratio of intermediates (excluding C6 ),
reservoir temperature and molecular weight of C7+ [30]. Correlation () is the proposed
correlation by Firoozabadi and Khalid.
 
CC2 − CC5
M M P = 9433 − 188 × 103
MwC7+ T 2.5
 
CC2 − CC5 2
+ 1430 × 103 (2.45)
MwC7+ T 2.5

Kuo Correlation

Kuo presented a correlation based on MMP prediction of four reservoir fluids and
several enriched-gas. This correlation can be used for prediction of methane concen-
tration at a given operating pressure or MMP for a fixed slug composition [31]. The
proposed correlation is as follow:
2 PVT of Gas Injection 51

Table 2.11 Constants of Kuo


A B C
correlation
0.19899861 0.00055769 0.58347828
D E F
0.62406453 0.57821035 0.00058948

log(C1 ) = (A + B × T ) × log(T ) + C × log(P)


+ D × log(MwC5+ )
+ (E + F × MwC2+ ) × log(MwC2+ ) (2.46)

where,
C1 : maximum allowable methane concentration in the injection gas, mol%
T: Temperature, °F.
P: pressure, psia.
MWC5+ : molecular weight of C5+ fraction in the reservoir fluid.
MWC2+ : molecular weight of intermediate (C2 –C4 ) fractions in the displacing
gas.
Table 2.11 represent the constant for Kuo’s correlation.

Pedrood Correlation

Pedrood simulated the rich gas injection and displacement process which is
investigated in one dimension and compositional model [32].

PM = 49.15 − 0.6863θ + 2.482 × 10−4 θ 2 − 0.2054 2 (2.47)

where  and θ present the properties of reservoir oil and injected gas respectively
and mathematically show as:

106 yC2−4
= (2.48)
[MwC5+ + (1.8T − 460)]

θ = 100(yC4 + 0.8yC3 + 0.5yC2 +C O2 ) (2.49)

2.4.6.3 The Impact of Optimization on the Accuracy of Correlations

As seen, the MMP correlations contain parameters which can be optimized for
specific systems. For instance, Zirrahi et al. used a large set of published data to
optimize parameters of Alston et al. [21] correlation to fit the measured MMP data
52 R. Azin et al.

for acid gas and flue gas streams [33]. They showed that a classic correlation with
optimized parameters works better than existing correlations.

2.4.7 Ternary Diagram

It is a common practice to study the miscibility concept using the ternary diagram.
Basically, ternary diagram is constructed for systems containing three components,
each laid on a corner of an equilateral triangle. In the case of a multi-component
hydrocarbon system, the whole mixture is divided into three subgroups or “pseudo”
components, usually named as light (L), intermediate (I), and heavy (H), as shown in
Fig. 2.12. The composition of reservoir oil and injection gas can be shown as a “point”
either on a side or inside the triangle. In Fig. 2.12 the example fluid includes 40%
component A, 40% component B, and 20% component C. The overall composition
can be easily divided into three parts based on defined L, I, and H components. These
components may be either pure or a combination of two or more pure components. As
an example, light components like CH4 , N2 and their mixture are normally regarded
as light; CO2 , H2 S, C2 –C6 as intermediate, and C7+ as the heavy group. A ternary
diagram may have two or more two-phase regions which represent the degree of
miscibility between these pseudo components.
It should be mentioned that any ternary diagram is constructed for a given temper-
ature and pressure. Changing either temperature or pressure causes the two-phase
region to expand or shrink and make the injection gas and reservoir oil approach or
deviate from miscibility, as well as a change in the slope of tie lines that determine
equilibrium compositions.

Fig. 2.12 A typical ternary


diagram
2 PVT of Gas Injection 53

Ternary diagram may be developed either by experiment or using EOS and PVT
modeling for a given fluid system. The question which arises in the PVT study of gas
injection is to what extent ternary diagram can represent the real behavior of reservoir
fluid and injection gas. As stated by Danesh [9], ternary diagrams and the procedure
of determining miscibility conditions between injection gas and reservoir oil using
ternary diagram should not be used for a real field operation and gas injection design.
In other words, the composition of reservoir fluid is so complex that reducing it into
simple subgroups of light, intermediate and heavy pseudo components may overlook
its real behavior when contacted injection gas. In field practice, ternary diagrams are
used in the preliminary design of gas composition enrichment for miscible injection,
which will then be overdesigned to guarantee the miscible conditions throughout the
gas flow in well and reservoir.

2.5 The Impact of Gas Injection on Phase Behavior


of Hydrocarbons

A case study of gas injection was done in this section. Gas cap was assumed as the
injection place. The changes in phase behavior of gas cap were studied first. Then, the
effect of new composition of gas cap and reservoir oil is studied. 1, 10, 30 years of gas
injection were considered as the injection time. CO2 , CH4 , N2 were selected as the
gas injection candidates. Fluid data of an Iranian oil reservoir are used in this case
study. Original gas cap composition was selected from the DL test of the studied
reservoir fluid. The original oil composition and reservoir pressure were assumed
constant during the gas injection process. Table 2.12 represents the gas cap and oil
composition, the initial volume of gas cap and gas injection rate into the gas cap in
all cases.
The procedure of calculation and modelling the phase behavior is as follows:
Step 1. Plot the P-T diagram for original oil and gas cap composition.
Step 2. Calculate the composition change in gas cap due to the gas injection using
the volume balance.
(VOrig × yi−Orig ) + (VI n j × yi−in j )
yi−new = (2.50)
VOrig + VI n j

where
yi = molar fraction of ith Component.
V = Volume (m3 ).
Step 3. Plot the P-T diagram for new gas cap composition.
Step 4. Specify the primary (oil) and secondary (gas) composition for evaluating
the phase behavior of the oil and gas mixture.
Step 5. Calculate the mole fraction of oil and gas in gas cap.
54 R. Azin et al.

Table 2.12 Composition of


Components Gas cap composition Oil composition
gas in the gas cap and
reservoir oil, initial volume of H2 S 0.513 0.63
gas cap and injection rate N2 3.2946 1.85
CO2 2.21 1.87
CH4 81.69 49.07
C2 H6 5.7649 5.82
C3 H8 2.68 4.1
IC4 0.62 1.21
NC4 1.2075 2.68
IC5 0.42 0.74
NC5 0.49 0.65
FC6 0.46 4.07
FC7 0.34 3.62
FC8 0.29 2.05
C09 7.05E−03 2.78
C10 4.56E−03 1.85
C11 2.96E−03 1.53
C12+ 5.43E−03 15.48
Initial gas cap volume (m3 ) Injection rate (m3 /day)
3,936,041,676 1,000,000

Step 6. Repeat steps 2–5 for the specified injection times.


The results of phase behavior modelling are presented in the following sections.

2.5.1 N2 Injection

Figure 2.13 shows the effects of Nitrogen injection phase behavior of gas in gas cap
and reservoir oil. As indicated in this figure, the saturation pressure of gas in gas
cap at reservoir temperature increases due to increase the amount of Nitrogen. Also,
Nitrogen injection decreases the criconder-therm and increases the criconder-bar of
the gas in gas cap. It can be understood that Nitrogen has significant effect on the
fluid saturation pressure.
2 PVT of Gas Injection 55

7000
Original Oil
Critical Point- Original
Original Gas Cap
6000 Mixture Composition- 1 year
Critical Point-Mixture 1 year
Gas Cap Composition- 1 year
Mixture Composition- 10 years
5000 Critical Point-Mixture 10 years
Gas Cap Composition- 10 years
Pressure (psi)

Mixture Composition- 30 years


Critical Point-Mixture 30 years
4000 Gas Cap Composition- 30 years

3000

2000

1000

0
-100 100 300 500 700 900 1100 1300
Temperature (F)

Fig. 2.13 Phase behavior of gas cap and reservoir oil for different times of Nitrogen injection

2.5.2 CO2 Injection

Figure 2.14 describe the changes of fluid phase behavior due to CO2 injection.
In opposite of Nitrogen injection, CO2 injection decreases the saturation pressure,
criconder-bar and criconder-therm of the gas in gas cap. It cannot be seen significant
changes in the reservoir fluid behavior by increasing the time of CO2 injection.

7000
Original Oil
Critical Point- Original
Original Gas Cap
6000 Mixture Composition- 1 year
Critical Point-Mixture 1 year
Gas Cap Composition- 1 year
Mixture Composition- 10 years
5000 Critical Point-Mixture 10 years
Pressure (psi)

Gas Cap Composition- 10 years


Mixture Composition- 30 years
Critical Point-Mixture 30 years
4000 Gas Cap Composition- 30 years

3000

2000

1000

0
-100 100 300 500 700 900 1100 1300
Temperature (F)

Fig. 2.14 Phase behavior of gas cap and reservoir oil for different times of CO2 injection
56 R. Azin et al.

2.5.3 CH4 Injection

The phase behavior of gas in gas cap and oil in methane injection are shown in
Fig. 2.15. This figure indicates that the methane has not a significant effect in gas cap
phase behavior in comparison of N2 and CO2 injection. But, the saturation pressure
of reservoir fluid increases more than CO2 injection and less than N2 injection.

2.6 Case Study of the Minimum Miscibility Pressure


for Different Injected Gases

A case study was conducted to determine the minimum miscibility pressure (MMP)
of the FCM and MCM processes for different injected gases. The oil sample was
introduced in Table 2.12. First, the minimum miscibility pressure was calculated
for the pure gas injection. Then, the effects of combining various injected gases
were investigated. In this case study, cell to cell method is applied for calculation of
MMP. This method determines the MMP by forming a ternary diagram. Figure 2.16
presents ternary diagrams for the mentioned oil sample and a mixture of CO2 and
CH4 as solvent. N2 and CH4 are categorized as pseudo component 3, C12+ as the
pseudo component 1 and other components as pseudo component 2. This diagram
shows the phase boundaries of combined fluid under various pressures at a constant
temperature. As it is obvious, the two-phase region, defined by the area in-between
the vapor and liquid lines, begins to close and decrease in size as pressure increases.
The miscibility condition is closed to the FCM by decreasing the size of the two

7000
Original Oil
Critical Point- Original
Original Gas Cap
6000 Mixture Composition- 1 year
Critical Point-Mixture 1 year
Gas Cap Composition- 1 year
Mixture Composition- 10 years
5000 Critical Point-Mixture 10 years
Gas Cap Composition- 10 years
Pressure (psi)

Mixture Composition- 30 years


Critical Point-Mixture 30 years
4000 Gas Cap Composition- 30 years

3000

2000

1000

0
-100 100 300 500 700 900 1100 1300
Temperature (F)

Fig. 2.15 Phase behavior of gas cap and reservoir oil for different times of Methane injection
2 PVT of Gas Injection 57

Fig. 2.16 A sample of ternary diagram for oil sample and a mixture of CO2 and CH4 as solvent in
different pressures

phase region. Also, the non-miscible condition is changed to the MCM as pressure
increases.
Figure 2.17 shows the MMP for pure injected gas in FCM and MCM mecha-
nisms. This figure demonstrates that the Nitrogen injection has the highest MMP in
both FCM and MCM. Also, the use of Methane as the injected gas need very high
pressure to reach the miscibility condition with the studied fluid. CO2 and Ethane

30000

FCM
25000
MCM
Minimum Miscibility Pressure (psi)

20000

15000

10000

5000

0
CH 4 CO 2 N2 C 2 H6
Injected Gas

Fig. 2.17 MMP for pure injected gas


58 R. Azin et al.

Minimum Miscibility Pressure (psi) 25000

20000 FCM
MCM

15000

10000

5000

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
CH4- Fraction

Fig. 2.18 MMP for CH4 -CO2 solvent

needs low pressure to meet the miscibility condition which can be reached in the
reservoir pressure and temperature. Also, it was found that Methane and Nitrogen
reach miscibility condition by the Vaporizing Gas Drive (VGD). In contrast, the
Condensing Gas Drive (CGD) is the main mechanism in CO2 and Ethane injection.

2.6.1 CH4 /CO2 Solvent

Figure 2.18 demonstrates the changes in MMP by increasing the fraction of methane
in injection of CH4 /CO2 solvent. As shown, the MMP would increase by increasing
the methane fraction in injected in fractions lower than 0.7 in both MCM and FCM
processes. The MMP of the MCM reaches a constant value and MMP of FCM
increase rapidly in the fraction upper than 0.7. Also, the CGD changes to VGD in
the 0.7 fraction of methane in injected gas.

2.6.2 CO2 /N2 Solvent

Figure 2.19 shows the MMP for FCM and MCM processes for CO2 and Nitrogen
solvent injection as a function of CO2 fraction in the solvent. The MMP of MCM
process in this case has a constant value until it reaches 0.5 fraction of CO2 in the
solvent. Also, in this value, the mechanism changes from VGD to CGD which cause
the MMP of MCM decreases from this value. The slope of decreasing the MMP of
FCM is lower in the higher fractions of CO2 .
2 PVT of Gas Injection 59

Minimum Miscibility Pressure (psi) 30000

25000
FCM
MCM
20000

15000

10000

5000

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
CO2- Mole Fraction

Fig. 2.19 MMP for CO2 -N2 solvent

2.6.3 CH4 /N2 Solvent

Figure 2.20 shows the MMP for Nitrogen and Methane solvent. In this case, the VGD
is the mechanism for reaching the miscibility condition. The MMP for MCM process
has a constant value due to the values of MMPs for pure Nitrogen and Methane which
almost similar as it is shown in Fig. 2.17.

30000

25000
Minimum Miscibility Pressure (psi)

FCM
MCM
20000

15000

10000

5000

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
N2- Mole Fraction

Fig. 2.20 MMP for N2 -CH4 solvent


60 R. Azin et al.

Minimum Miscibility Pressure (psi) 25000

20000 FCM
MCM

15000

10000

5000

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
CH4- Fraction

Fig. 2.21 MMP for C2 H6 -CH4 solvent

2.6.4 CH4 /C2 H6 Solvent

Figure 2.21 show the MMP for Ethane and Methane solvent as a function of Methane
fraction. As it is shown, the MMP increases by increasing the fraction of Methane in
the solvent in both FCM and MCM mechanisms. Also, it was found that the fraction
0.2 is the threshold for changing the CGD to VGD. This change causes that the
MMP of FCM increased rapidly but, the MMP for MCM process increased slowly
in a higher amount of Methane in the solvent.

2.6.5 C2 H6 /CO2 Solvent

Figure 2.22 describes the MMP of the Ethane and CO2 solvent as a function of CO2
fraction. Obviously, the MMP of FCM has constant value due to the close MMP of
pure CO2 and Ethane. But, the MMP of MCM increases slightly by increasing the
CO2 fraction in the solvent. Also, the mechanism of reaching miscibility in this case
is CGD.

2.6.6 C2 H6 /N2 Solvent

Figure 2.23 shows the MMP for Ethane and Nitrogen solvent in FCM and MCM
process. The MMP of FCM process increases by increasing the Nitrogen fraction.
2 PVT of Gas Injection 61

5000
FCM
4500
Minimum Miscibility Pressure (psi)

MCM
4000

3500

3000

2500

2000

1500

1000

500

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
CO2- Fraction

Fig. 2.22 MMP for C2 H6 -CO2 solvent

30000
Minimum Miscibility Pressure (psi)

25000
FCM
MCM
20000

15000

10000

5000

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
N2- Fraction

Fig. 2.23 MMP for C2 H6 -N2 solvent

Also, the MMP of MCM process increases until the fraction of Nitrogen reaches 0.5.
But, the MMP becomes constant in the higher fractions of Nitrogen. It is found out
that in the 0.2 fraction of Nitrogen CGD turns into the VGD. This fact shows that
the Nitrogen has higher power to control the miscibility process.
62 R. Azin et al.

2.7 The Effect of Gas Injection on Asphaltene Precipitation

In this section, a case study was conducted on asphaltene precipitation due to injection
of different gases into an Iranian oil reservoir [34]. For this purpose, the composition
of reservoir fluid is split and regrouped. Then, the new heavy component (C36+ ) is
divided into precipitating and non-precipitating components. After that, the equation
of state is tuned using the experimental data of the studied fluid. Tables 2.13 and
2.14 show the original composition and modified composition of reservoir fluid.
Table 2.15 shows the SARA analysis of the studied fluid.
Multi-phase flash can model the asphaltene precipitation and the solid model can
predict the fugacities of components in solid phase. The solid phase can contain
one or more components. The precipitated phase is assumed as an ideal mixture of
solid components. Precipitating component fugacity is calculated by the following
relationship:

Table 2.13 Reservoir oil


Components Reservoir oil
composition and heavy
(Mol %)
fraction properties
H2 S 0.27
N2 0.71
CO2 1.55
C1 40.01
C2 8.12
C3 5.35
iC4 1.02
nC4 2.45
iC5 1.04
nC5 1.32
C6 4.94
C7 4.67
C8 2.92
C9 3.79
C10 2.95
C11 2.75
C12 + 16.15
Total 100.00
Molecular weight of residual oil 285
+
Molecular weight of C12 fraction 548
Molecular weight of reservoir oil 128
Sp.Gr. of C12 + fraction @ 60 °F 0.9286
2 PVT of Gas Injection 63

Table 2.14 Reservoir oil


Components Reservoir oil (Mol %)
composition after splitting
and regrouping N2 0.2699678
CO2 0.70991534
H2 S 1.5498152
CH4 40.005229
C2 H6 8.1190317
C3 H8 5.349362
IC4 1.0198784
NC4 2.4497078
IC5 1.039876
NC5 1.3198426
C6 –C11 22.017374
C12 –C20 4.4833402
C21 –C29 3.2387371
C30 –C35 1.6426098
C36A+ 6.3776687
C36B+ 0.4076444

Table 2.15 Results of SARA


Saturates (wt. Aromatics Resins (wt%) Asphaltenes
analysis
%) (wt%) (wt%)
66.70 29.80 0.70 2.80

vs ( p − p ∗ )
ln f s = ln f s∗ + (2.51)
RT

where fs * and fs are the fugacities of pure solid asphaltenes at pressure p* and p,
respectively. vs is the molar volume of pure solid asphaltenes. R and T are the
universal gas constant and temperature, respectively.
The mole fraction of C36A+ and C36B+ is computed as follow:

wC36B+ × MwOil
z C36B+ = (2.52)
MwC36B+

z C36A+ = z C36+ − z C36B+ (2.53)

After these calculations, the asphaltene precipitation is predicted by defining a


secondary composition. Five cases were selected as composition of injection gas
including, N2 , CO2 , CH4 , and two mixtures of gases (one CH4 dominant and one
CO2 dominant). In the following sections, the results of the asphaltene precipitation
are described.
64 R. Azin et al.

0.5 Solvent Mole Fraction = 0.00


0.45 Solvent Mole Fraction = 0.05
Solvent Mole Fraction = 0.1
Asphaltene Precipitation (wt%)

0.4
Solvent Mole Fraction = 0.15
0.35 Solvent Mole Fraction = 0.2

0.3

0.25

0.2

0.15

0.1

0.05

0
0 2000 4000 6000 8000 10000 12000 14000
Pressure (psi)

Fig. 2.24 Curve of asphaltene precipitation due to Nitrogen injection

2.7.1 Pure N2

Figure 2.24 represents the impact of pure N2 injection on asphaltene precipitation


which is investigated on 0, 0.05, 0.1, 0.15 and 0.2 mol fraction of N2 . It is shown that
the maximum amount of asphaltene precipitation is increased slightly by increasing
the mole fraction of injected N2 . Also, the pressure range of asphaltene precipita-
tion is increased due to the presence of N2 . This figure indicates that by increasing
the mole percent of N2 , the maximum amount of asphaltene precipitation can be
reached in a higher pressure in N2 injection scheme. As the maximum amount of
asphaltene precipitation occurs in the saturation pressure, it can be concluded that
N2 can significantly increase the saturation pressure.

2.7.2 Pure CO2

The injection of pure CO2 increases the maximum asphaltene precipitation greatly, as
shown in Fig. 2.25. The onset pressure of asphaltene precipitation is also increased
due to increasing the mole percent of CO2 . In this scheme, the pressure range of
asphaltene precipitation increases due to increasing the onset pressure. It can be
concluded that the CO2 effect on the saturation pressure is lower than N2 injection.
2 PVT of Gas Injection 65

0.8
Solvent Mole Fraction = 0.00
0.7 Solvent Mole Fraction = 0.05
Asphaltene Precipitation (wt%)

Solvent Mole Fraction = 0.1


0.6 Solvent Mole Fraction = 0.15
Solvent Mole Fraction = 0.2
0.5

0.4

0.3

0.2

0.1

0
0 2000 4000 6000 8000 10000 12000 14000
Pressure (psi)

Fig. 2.25 Curve of asphaltene precipitation due to CO2 injection

2.7.3 Pure CH4

Figure 2.26 represents the effect of pure CH4 injection on the asphaltene precipitation.
The impact of pure CH4 on the maximum amount of asphaltene precipitation is
similar to N2 . This amount increases slightly in this scheme. Also, the onset pressure
of asphaltene precipitation increases slightly in pure CH4 injection in comparison to

0.5 Solvent Mole Fraction = 0.00


0.45 Solvent Mole Fraction = 0.05
Asphaltene Precipitation (wt%)

Solvent Mole Fraction = 0.1


0.4
Solvent Mole Fraction = 0.15
0.35 Solvent Mole Fraction = 0.2

0.3

0.25

0.2

0.15

0.1

0.05

0
0 2000 4000 6000 8000 10000 12000 14000
Pressure (psi)

Fig. 2.26 Curve of asphaltene precipitation due to Methane injection


66 R. Azin et al.

0.5
Solvent Mole Fraction = 0.00
0.45 Solvent Mole Fraction = 0.05
Asphaltene Precipitation (wt%)

0.4 Solvent Mole Fraction = 0.1


Solvent Mole Fraction = 0.15
0.35 Solvent Mole Fraction = 0.2
0.3

0.25

0.2

0.15

0.1

0.05

0
0 2000 4000 6000 8000 10000 12000 14000

Pressure (psi)

Fig. 2.27 Curve of asphaltene precipitation due to Methane dominant gas injection

N2 and CO2 . The impact of CH4 on the saturation pressure is higher than CO2 and
lower than N2 .

2.7.4 CH4 Dominant

In this scheme, the mole percent of CH4 , CO2 and H2 S are 0.9, 0.05 and 0.05,
respectively. Figure 2.27 shows the effect of this scheme of gas injection on asphaltene
precipitation. This effect is almost similar to CH4 injection due to high mole fraction
of methane. But, the maximum of asphaltene precipitation and onset pressure are
lower in this scheme in comparison to the pure methane injection.

2.7.5 CO2 Dominant

The mole fraction of CO2 , CH4 and H2 S are 0.8, 0.1 and 0.1, respectively. As it is
shown in Fig. 2.28, the maximum amount of asphaltene precipitation and the onset
pressure are lower than pure CO2 injection scheme which is the impact of impurity
in reducing the effect of pure CO2 on asphaltene precipitation.
2 PVT of Gas Injection 67

0.8
Solvent Mole Fraction = 0.00
0.7 Solvent Mole Fraction = 0.05
Asphaltene Precipitation (wt%)

Solvent Mole Fraction = 0.1


0.6 Solvent Mole Fraction = 0.15
Solvent Mole Fraction = 0.2
0.5

0.4

0.3

0.2

0.1

0
0 2000 4000 6000 8000 10000 12000 14000
Pressure (psi)

Fig. 2.28 Curve of asphaltene precipitation due to CO2 dominant gas injection

2.7.6 Comparison of Different Gases

Figure 2.29 comprises the effect of 0.1 mol fraction of different injection gases on
asphaltene precipitation in reservoir temperature (174 °F). As it obvious, CO2 injec-
tion increases the maximum amount of precipitated asphaltene in comparison of
other injection gases. Also, the onset pressure of asphaltene precipitation increases
in the case of CO2 injection. In CO2 dominant gas injection, the maximum asphal-
tene precipitation decreased due to the presence of CH4 and H2 S. The lowest onset

0.65
0.6 CH 4

0.55 CO 2
0.5 N2
Asphaltene Precipitation (wt%)

0.45
CO 2 Dominant
0.4
CH 4 Dominant
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000
Pressure (psi)

Fig. 2.29 Curve of asphaltene precipitation for 0.1 mol fraction of different injected gas
68 R. Azin et al.

pressure belongs to CH4 dominant scheme. Also, it can be concluded, that the best
choice for gas injection is CH4 dominant for preventing a high amount of asphaltene
precipitation and a significant increase in saturation pressure.

2.8 Optimum Design of Gas Injection

Prediction of miscibility conditions includes MMP determination and optimum gas


composition for miscible injection into a given oil reservoir. The latter is known
as minimum miscibility enrichment (MME) which refers to gas composition. It
is worth noting that MMP is sometimes known as minimum dynamic miscibility
pressure (MDMP) [20]. This term is used to reflect the dynamic behavior of gas oil
contact and gradual enrichment which eventually leads to miscibility conditions. It
is well known that miscible gas injection works best and gives highest ultimate oil
recovery factor (RF). Gases like CO2 and H2 S lower the MMP upon contact with
reservoir oil, while light components like N2 and CH4 increase the mixture MMP [33].
Similar trend was also observed for oxygen, and it was found that this component
increases MMP when present in CO2 stream, but not as much as when CO2 stream is
contaminated with N2 [35]. Therefore, although co-injection of contaminants such as
CH4 , H2 S, N2 , O2 and SO2 with CO2 stream as acid gas and flue gas provides an extra
benefit to operators and let them handle greenhouse gas emission from reservoirs or
plants in addition to miscible EOR and CO2 sequestration projects, each individual
contaminant may increase or decrease MMP. The combined effect of components
depends on composition of injection gas and needs be examined and modeled.

2.9 PVT Challenges Associated with Gas Injection

From what we have presented here, PVT studies play a crucial role in the design
and optimization of the gas injection process. However, there are challenges associ-
ated with PVT, part of which is reflected in fluid characterization, QC of the reser-
voir and injection fluids, as discussed above. In addition, there are challenges with
constructing and using ternary diagram when there is a risk of asphaltene precip-
itation and deposition during gas injection, as it alters the whole composition and
properties of reservoir oil when contacts the injected gas.
The dynamic behavior of phase envelope may be simulated through PVT experi-
ments. Results of these tests can be used to determine the method of secondary/tertiary
recovery. For example, if differential liberation (DL) test reveals a shift of phase
diagram to higher temperature, the fluid composition becomes heavier as a result of
light hydrocarbon removal as vapor phase. Such a change in composition acts as a
drive toward higher liquid condensation, and the reservoir is a good candidate for
either pressure maintenance by lean gas injection to keep intermediate hydrocarbons
or for gas cycling to vaporize and recover liquid hydrocarbons [36]. Orr and Taber
2 PVT of Gas Injection 69

describe more suggestions for secondary/tertiary recovery techniques based on phase


behavior of oil reservoirs.
Usually, PVT experiments are performed on bulk of fluid, whereas the reservoir
fluid exists in porous media. As long as the pores are in the order of micrometers
or larger, the measured PVT properties at bulk conditions are close to the porous
medium. For smaller pores in the order of micrometers to nanometers, the pore
size affects PVT properties. Such small sizes eventuate considerable curvature effect
which results in significant capillary forces [37]. The effects of these forces are in
opposite to gravity such that instead of a sharp change in phase, there will be a
transition zone with two or three phases in equilibrium with each other [38]. The
capillary effect is remarkable in formations with pore radius of micro and nano-
orders and high interfacial tension. One of the important outcomes of this fluid
confinement is variation in phase behavior. A detailed review and analysis of the
effect of capillary pressure on phase behavior studies of gas condensate systems is
given by Kiani et al. [39]. They found that presence of porous media has significant
effects on phase behavior and VLE calculations. These effects could be due to shift
in critical properties and also difference between phase pressures related to capillary
forces. Their results indicate that interfacial curvature will suppress bubble point
pressure; therefore, molar fraction of light components will decrease, while heavy
components will increase in their composition. Because of low order of magnitude
for C12+ composition, increase in composition is not significant. Also, gas phase
pressure will decrease. It is investigated that porous media has a remarkable effect
on GOC location, as pore radius decreases, bubble point pressure decreases too,
resulted to change phases in lower depths. With an increase in pore sizes, decrease
in compositional gradients occurs. Due to suppression in bubble point pressure and
GOC appearance in upper depths, the thickness of the transition zone decreases and
consequently, compositional gradients calculated in this zone decreases.
Asphaltene deposition is another challenge associated with PVT of gas injec-
tion, which in turn affects reservoir petro-physical properties and alters wetting
conditions. This subject is studied in experimental, numerical and field level. For
example, Mousavi-Dehghani et al. studied asphaltene deposition in porous media
during a miscible gas injection [40]. They observed a decrease in porosity of porous
medium upon gas injection and interpreted this by asphaltene deposition as a result
of changes in oil composition upon gas injection. Change in reservoir permeability
due to asphaltene precipitation in core samples put under gas injection is reported
by Minssieux et al. [41]. They suggested a mechanism for core damage by asphal-
tene deposition which describes rate of asphaltene separation and deposition on rock
surface. The wells in Ghawar oil field, Saudi Arabia’s main oil field, have also faced
with asphaltene problem upon gas injection [42]. In Iran, asphaltene deposition is
reported in a number of gas injection projects, including Kupal [3], and Aghajari
[43]. A detailed case study on the impact of gas injection on asphaltene precipitation
was presented in Sect. 2.7.
Another issue involves scale up of experimental data for miscibility measure-
ments. For example, there are debates on the applicability of MMP measurements
obtained from slim tube. This apparatus is made of an unconsolidated sand pack to
70 R. Azin et al.

mimic porous media in a capillary tube. However, it is too simple to represent a real
reservoir and quite far from a real porous medium. Therefore, the oil recovery factor
obtained from the slim tube has little or no correlation with that obtained in reservoir.
In addition, the slim tube experiment is conducted at “local” rather than “overall”
thermodynamic equilibrium between reservoir oil and injection gas. Therefore, slim
tube can be considered as an “experimental simulation” of miscibility test.

References

1. Bennion DB, Thomas EB, Bennion DW, Bietz RE. Formation screening to minimize perme-
ability impairment associated with acid gas or sour gas injection/disposal. In: Annual technical
meeting. Petroleum Society of Canada; 1996.
2. Hustad OS, Jia N, Pedersen KS, Memon A, Leekumjorn S. High-pressure data and modeling
results for phase behavior and asphaltene onsets of Gulf of Mexico oil mixed with nitrogen.
SPE Reserv Eval Eng. 2014;17:384–95.
3. Kalantari-Dahaghi A, Gholami V, Moghadasi J, Abdi R. Formation damage through asphaltene
precipitation resulting from CO2 gas injection in Iranian carbonate reservoirs. SPE Prod Oper.
2008;23:210–4.
4. Yonebayashi H, Al Mutairi AM, Al Habshi AM, Urasaki D. Dynamic asphaltene behavior for
gas-injection risk analysis. SPE Reserv Eval Eng. 2011;14:493–504.
5. Negahban S, Bahamaish JNM, Joshi N, Nighswander J, Jamaluddin AKM. An experimental
study at an Abu Dhabi reservoir of asphaltene precipitation caused by gas injection. SPE Prod
Facil. 2005;20:115–25.
6. Srivastava RK, Huang SS, Dong M. Asphaltene deposition during CO2 flooding. SPE Prod
Facil. 1999;14:235–45.
7. Yonebayashi H, Masuzawa T, Dabbouk C, Urasaki D. Ready for gas injection: asphaltene
risk evaluation by mathematical modeling of asphaltene precipitation envelope (APE) with
integration of all laboratory deliverables. In: SPE reservoir characterisation and simulation
conference. European Association of Geoscientists & Engineers; 2009. p. cp-170.
8. Yonebayashi H, Takabayashi K, Iizuka R, Tosic S. Managing experimental-data shortfalls
for fair screening at concept selection: case study to estimate how acid-gas injection affects
asphaltene-precipitation behavior. Oil Gas Facil. 2016;5.
9. Danesh A. PVT and phase behaviour of petroleum reservoir fluids. Elsevier; 1998.
10. Osfouri S, Azin R. An overview of challenges and errors in sampling and recombination of
gas condensate fluids. J Oil Gas Petrochemical Technol. 2016;3:1–14.
11. Osfouri S, Azin R, Rezaei Z, Moshfeghian M. Integrated characterization and a tuning strategy
for the PVT analysis of representative fluids in a gas condensate reservoir. Iran J Oil Gas Sci
Technol. 2018;7:40–59.
12. Gerami S, Kiani Zakheradi M, Azin R, Osfouri S. A unified approach for quality control of
drilled stem test (DST) and PVT data. Gas Process. 2014;2:40–50.
13. Kesler MG, Lee BI. Improved prediction of enthalpy of fractions: hydrocarbon Processing;
1976.
14. Green DW, Willhite GP. Enhanced oil recovery, vol. 6. Henry L. Doherty Memorial Fund of
AIME, Society of Petroleum Engineers Richardson, TX; 1998.
15. Chen SM, Olynyk J, Asgarpour S. Effect of multiple-contact miscibility on solvent slug size
determination. J Can Pet Technol. 1986;25.
16. Clark NJ, Shearin HM, Schultz WP, Garms K, Moore JL. Miscible drive—its theory and
application. J Pet Technol. 1958;10:11–20.
17. Holm LW. Miscibility and miscible displacement. J Pet Technol. 1986;38:817–8.
2 PVT of Gas Injection 71

18. Gu Y, Hou P, Luo W. Effects of four important factors on the measured minimum miscibility
pressure and first-contact miscibility pressure. J Chem Eng Data. 2013;58:1361–70.
19. Yuan H, Johns RT, Egwuenu AM, Dindoruk B. Improved MMP correlations for CO2 floods
using analytical gas flooding theory. In: SPE/DOE symposium on improved oil recovery.
Society of Petroleum Engineers; 2004.
20. Johnson JP, Pollin JS. Measurement and correlation of CO2 miscibility pressures. In: SPE/DOE
enhanced oil recovery symposium. Society of Petroleum Engineers; 1981.
21. Alston RB, Kokolis GP, James CF. CO2 minimum miscibility pressure: a correlation for impure
CO2 streams and live oil systems. Soc Pet Eng J. 1985;25:268–74.
22. Cronquist C. Carbon dioxide dynamic miscibility with light reservoir oils. In: Proceedings of
the Fourth Annual US DOE Symposium, Tulsa, vol. 1; 1978. p. 28–30.
23. Shokir EME-M. Precise model for estimating CO2 -oil minimum miscibility pressure. Pet
Chem. 2007;47:368–76.
24. Shokir EME-M. CO2 -oil minimum miscibility pressure model for impure and pure CO2
streams. J Pet Sci Eng. 2007;58:173–85.
25. Emera MK, Sarma HK. A reliable correlation to predict the change in minimum miscibility
pressure when CO2 is diluted with other gases. SPE Reserv Eval Eng. 2006;9:366–73.
26. Kamari A, Arabloo M, Shokrollahi A, Gharagheizi F, Mohammadi AH. Rapid method to
estimate the minimum miscibility pressure (MMP) in live reservoir oil systems during CO2
flooding. Fuel. 2015;153:310–9.
27. Kaydani H, Najafzadeh M, Hajizadeh A. A new correlation for calculating carbon dioxide
minimum miscibility pressure based on multi-gene genetic programming. J Nat Gas Sci Eng.
2014;21:625–30.
28. Glaso O. Miscible displacement: recovery tests with nitrogen. SPE Reserv Eng. 1990;5:61–8.
29. Glaso O. Generalized minimum miscibility pressure correlation (includes associated papers
15845 and 16287). Soc Pet Eng J. 1985;25:927–34.
30. Firoozabadi A, Khalid A. Analysis and correlation of nitrogen and lean-gas miscibility pressure
(includes associated paper 16463). SPE Reserv Eng. 1986;1:575–82.
31. Kuo SS. Prediction of miscibility for the enriched-gas drive process. In: SPE annual technical
conference and exhibition. Society of Petroleum Engineers; 1985.
32. Pedrood P. Prediction of minimum miscibility pressure in rich gas injection. MSc thesis, Tehran
Univ Tehran; 1995.
33. Zirrahi M, Azin R, Malakooti R. Prediction of minimum miscibility pressure for carbon capture
and sequestration and acid gas disposal applications; n.d.
34. Izadpanahi A, Azin R, Osfouri S, Malakooti R. Asphaltene precipitation in a light oil reservoir
with high producing GOR: case study. Adv Nanomater Technol Energy Sect. 2019;3:270–9.
35. Yang F, Zhao G-B, Adidharma H, Towler B, Radosz M. Effect of oxygen on minimum
miscibility pressure in carbon dioxide flooding. Ind Eng Chem Res. 2007;46:1396–401.
36. Orr Jr FM, Taber JJ. Phase diagrams (1987 PEH Chapter 23). Pet Eng Handb. 1987.
37. Firoozabadi A. Thermodynamics of hydrocarbon reservoirs. McGraw-Hill; 1999.
38. Wheaton RJ. Treatment of variations of composition with depth in gas-condensate reservoirs
(includes associated papers 23549 and 24109). SPE Reserv Eng. 1991;6:239–44.
39. Kiani M. Modeling development of compositional grading in hydrocarbon reservoirs. Persian
Gulf University; 2019.
40. Mousavi DSA, Vafaei SM, Mirzaei B, Fasih M. Experimental investigation on asphaltene
deposition in porous media during miscible gas injection; 2007.
41. Minssieux L, Nabzar L, Chauveteau G, Longeron D, Bensalem R. Permeability damage due to
asphaltene deposition: experimental and modeling aspects. Rev l’Institut Français Du Pétrole.
1998;53:313–27.
42. Sunil K, Abdullah A-G, Dimitrios K. Asphaltene precipitation in high gas-oil ratio wells. In:
Middle east oil show; 2003.
43. Hashemi R, Kshirsagar LK, Nandi S, Jadhav PB, Golab EG. Experimental and CMG study
of asphaltene precipitation under natural depletion and gas injection conditions. Pet Coal.
2019;61.
72 R. Azin et al.

44. Izadpanahi A, Azin R, Osfouri S, Malakooti R. MODELING OF ASPHALTENE PRECIPI-


TATION IN A LIGHT OIL RESERVOIR WITH HIGH PRODUCING GOR: CASE STUDY.
In2nd International Biennial Oil, Gas and Petrochemical Conference 2018 Nov.
45. Izadpanahi A, Azin R, Osfouri S, Malakooti R. Optimization of Two Simultaneous Water and
Gas Injection Scenarios in a High GOR Iranian Oil Field. In82nd EAGE Annual Conference
& Exhibition 2020 Dec 8 (Vol. 2020, No. 1, pp. 1-5). European Association of Geoscientists
& Engineers.
Chapter 3
Basics of Oil and Gas Flow in Reservoirs

Reza Azin, Amin Izadpanahi, and Parviz Zahedizadeh

Abstract This chapter deals with basic concepts of oil and gas flow in porous media.
In the first section (oil flow), the basics of oil flow in porous media and different flow
equations, including viscous, Darcy, and Brinkman flow are introduced. Then, based
on a new work the boundary between these kinds of flow are reported. Then, the well-
known diffusivity equation, combination of continuity equation and Darcy velocity,
is introduced. The relative permeability as an important concept in two-phase flow
through porous media is described as well. After that, the forces affecting oil flow
in porous media are introduced as well as different dimensionless numbers which
reflect the relative importance of forces during oil flow. The gas flow through porous
reservoir can occur as single or multiphase. A second phase releases as a result of pres-
sure decline in reservoir and starts flowing when its saturation exceeds a minimum
value, known as critical saturation. At these conditions, the relative permeability
concept applies to take into account the multiphase flow. Also, the classic Darcy’s
law equation for flow in porous media may fail in certain flowing conditions, and
the non-Darcy flow equations need to be studied and applied for proper modelling of
gas flow in porous medium. In addition, the role of wettability alteration on produc-
tion enhancement of gas condensate reservoirs is a novel approach in gas reservoir
performance. Also, the active forces during CO2 storage and the challenges of gas
flow in unconventional gas reservoirs are two important subjects. These subjects will
be covered in the second section (gas flow) of this chapter.

R. Azin (B)
Faculty of Petroleum, Gas and Petrochemical Engineering, Department of Petroleum Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Izadpanahi · P. Zahedizadeh
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
P. Zahedizadeh
e-mail: P.Zahedizadeh@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 73


R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_3
74 R. Azin et al.

Nomenclature

Acronyms

ANN Artificial neural network


BHP Bottom-hole pressure, psi
IPR Inflow performance curve
LSSVM Least square support vector machine
MGGP Multi-gene genetic programming
TPR Tubing performance curve

Variables

C Isothermal compressibility, pa− 1


Ct Total compressibility, 1/psi
E Non-Darcy effect
Fo Forchheimer number
f Fractional flow
G Cumulative gas (HCPV)
Grw Diffusional Grashof number
h Formation thickness, m
I0 Modified Bessel function, first kind, zero-order
I1 Modified Bessel function, first kind, the first order
k Permeability, md
K0 Modified Bessel function, the second kind, zero-order
K1 Modified
 √ Bessel
 function, the second kind, the first order
M M = zf(z)
p Pressure, psi
p̄ Average reservoir pressure, pa
p0 Initial Pressure, psi
pp Pseudo-pressure
Pi Initial reservoir pressure, pa
Pe Peclet number
PD Dimensionless pressure, (PD = 2πKf h(Pqμi −P(r,t)) )
Pi Initial reservoir pressure, pa
Pwf Wellbore pressure during the drawdown period, pa
i −Pwf )
PDwf Dimensionless wellbore pressure, (PDwf = 2πKf h(P qμ
)
3 −1
q Production rate, m s
qD Dimensionless production, (qD = kf h(PqμB
i −Pwf)
)
R Cylinder Radius, m
Ra Rayleigh number
3 Basics of Oil and Gas Flow in Reservoirs 75

Re Reynolds number
r Radius, ft
rc Pore radius, ft
re Reservoir outer boundary radius, m
rw Wellbore radius, m
rD Dimensionless radius, (rD = r/rw )
rDe Dimensionless outer boundary radius, (reD = re /rw )
S A characteristic coefficient of the fractured rock proportional to the
specific surface of the block
Sc Schmidt number
t Time, s
tD Dimensionless time, (tD = [(∅C) +(∅C)
Kf t
]μr2 )
m f w
U Velocity, ms− 1
U* Rate of mass flow per unit volume, represents the transfer of fluid between
blocks and fractures, kg/s m3 , (U∗ = ρSKm (Pμm −Pf ) )
U Flux, S− 1
v Velocity, ft/s
v Average velocity (m/s)
vg General solution
vp Particular solution
z Compressibility factor/Laplace space variable

Greek Letters

μ Viscosity, cp
β Coefficient of quadratic velocity term
γ Coefficient of cubic velocity term
σ Interfacial tension, dyne/ft
ρ Density, lb/ft3
v Velocity (ft/d)
Ø Porosity
αK r2
λ Dimensionless matrix/fracture permeability ratio, (λ = K12 w )
ω Fracture storage, (ω = ∅1 C∅12+∅
C2
2 C2
)
θ Rock wettability
α Angle of deviation/interporosity flow shape factor, m− 2

Subscripts

c Capillary
D Dimensionless
76 R. Azin et al.

f Fracture
gde Dry gas
gw Wet gas
id Dry gas injected
m Matrix
pD Recovery of wet gas
r Relative
W Wellbore
wc Critical water

3.1 Introduction

This chapter deals with basic concepts of oil and gas flow in porous media. In the
first section (oil flow), the basics of oil flow in porous media and different flow
equations, including viscous, Darcy, and Brinkman flow are introduced. Then, based
on a new work the boundary between these kinds of flow are reported. Then, the well-
known diffusivity equation, combination of continuity equation and Darcy velocity,
is introduced. The relative permeability as an important concept in two-phase flow
through porous media is described as well. After that, the forces affecting oil flow
in porous media are introduced as well as different dimensionless numbers which
reflect the relative importance of forces during oil flow.
The gas flow through porous reservoir can occur as single or multiphase. A second
phase releases as a result of pressure decline in reservoir and starts flowing when its
saturation exceeds a minimum value, known as critical saturation. At these condi-
tions, the relative permeability concept applies to take into account the multiphase
flow. Also, the classic Darcy’s law equation for flow in porous media may fail in
certain flowing conditions, and the non-Darcy flow equations need to be studied and
applied for proper modelling of gas flow in porous medium. In addition, the role
of wettability alteration on production enhancement of gas condensate reservoirs
is a novel approach in gas reservoir performance. Also, the active forces during
CO2 storage and the challenges of gas flow in unconventional gas reservoirs are two
important subjects. These subjects will be covered in the second section (gas flow)
of this chapter.

3.2 Oil Flow

3.2.1 Basics of Oil Flow in Porous Media

The original form of Navier–Stokes equation, (shown in Eq. (3.1)) describes the
balance of pressure forces, viscous forces, and external forces. A specific solution
3 Basics of Oil and Gas Flow in Reservoirs 77

of Navier–Stokes equation is presented in Eq. (3.2). The assumptions behind the


Navier–Stokes equation include incompressible fluid with constant fluid properties
and Newtonian fluid. According to the equation of motion, the rate of increase of
momentum equals the sum of rate of momentum addition by convection, molecular
transport, and external forces on fluid, all expressed per unit volume [1]. These are
the main forces acting in a flowing continuum. Flow of fluids through porous media
may be treated by simplifying the porous media as a bundle of capillary tubes. The
Poiseuille equation, also known as Hagen–Poiseuille equation, shown in Eq. (3.3) and
is used to formulate the flow of fluids in simplified porous medium. Also, fluid flow
through porous media was studied experimentally by Darcy on an unconsolidated
pack of sandstone [2].

∂ 2v ∂ 2v 1 dp
+ = (3.1)
∂x2 ∂ y2 μ dz
1 dp  2 
v= R − r2 (3.2)
4μ dz

1 R 2 dp
v= (3.3)
8 μ dz

He constructed a vertical physical model, packed it with unconsolidated sand,


and saturated the medium with water. Then, he performed experiments at atmo-
spheric conditions to derive his famous equation, now widely used in the oil industry
(Eq. (3.4)).

K dP
v= (3.4)
μ dz

He introduced the term “permeability” to the context of fluid flow through porous
media. The assumptions made by Darcy referred to the physical model in which
he conducted his experiments. These assumptions include incompressible fluid,
steady state conditions, complete saturation, single phase flow, constant physical
properties (mainly viscosity and density), laminar flow, homogeneous formation,
and constant temperature. This equation is now widely applied in all types of fluid
flow through porous media, including groundwater flow, oil flow, and gas flow. By
equating Poiseulle and Darcy equations, an expression for permeability is obtained
as a function of capillary radius. The Poiseulle equation can be written for both
pressure gradient and gravity flow [3]. The Darcy velocity is sometimes referred to
infiltration velocity [4]. The Darcy equation was later extended to conditions which
were not originally considered by Darcy. For example, cases like unsteady state
and pseudosteady state flow of slightly compressible and compressible fluids, as
well as turbulent flow and multiphase flow in heterogeneous and layered formations
quite deviate from original assumptions made by Darcy. However, the researchers
and petroleum engineers preferred to keep the original form of Darcy’s equation
78 R. Azin et al.

and modify it whenever necessary and extend its application as the foundation for
oil flow in reservoir. The Darcy’s law equation can be expressed in both pressure
gradient and potential gradient forms to account for gravity effects and vertical flow
of fluids [5]. Though derived from experiments, the Darcy’s law may be derived
theoretically from the Navier–Stokes equation using a formal averaging procedure
[6]. According to Saidi [3, 7], Darcy equation can be derived by ignoring the inertia
terms in the Navier–Stokes equation and taking the average velocity of the flow.
Also, Neuman showed that the Hagen-Poiseuille equation, as well as the expression
describing Couette flow between parallel plates, can be derived from the equations
presented in his work and may thus be viewed as special cases of Darcy’s law [6].
The Darcy flow is sometimes referred to slip flow. This means that flow on the solid
walls, here on grain surfaces of porous medium, is not zero, and the fluid molecules
slip on the solid walls. This is the main difference between Darcy flow and viscous
flow governed by Navier–Stokes equation. The viscous flow stops when the radius of
a capillary tube (or capillary pore) is close to the thickness of adsorbed layer. Under
these conditions, slip flow prevails and the viscous flow is no longer valid. When the
pore diameter is small enough to fit within the boundary layer of viscous flow, the
fluid will be unable to flow under viscous regime, and slip flow appears. Darcy flow
occurs when slip flow fully develops. Saidi states that flow through a circular tube
with a diameter of about one micron is probably non-viscous and probably follows
Darcy Law [3, 7].
In recent years, there have been debates and questions on the applicability and
validity of Darcy’s law in low-permeable formations, where fluid velocity is typi-
cally low. Many researches addressed several limitations of Darcy’s law at certain
conditions where this equation is no longer valid [8–11]. These conditions are high
velocities, known as non-Darcy flow in the context of petroleum engineering, low gas
pressure, known as slip flow or Klinkenberg effect [12], and Non-Newtonian flow as
observed in certain heavy oil, which is addressed by Gavin [9]. He analyzed typical
velocities in oil and gas reservoirs and realized significant differences from an order
of magnitude or more, between these velocities and those used in original Darcy’s
experiments [2]. In all oil and gas cases, the flow velocity was much lower than the
ranges of velocities reported by Darcy. Going farther from producing well reduces
the oil and gas velocity so that there is little difference between linear and non-linear
solutions of diffusivity equation [13]. Longmuir mentioned that the pre-Darcy flow
may occur both in consolidated, tight systems, and in unconsolidated sand packs,
and came to a conclusion that there is lack of accessible data which confirms Darcy’s
law governs flow of fluid in porous media at reservoir conditions [9]. When fluid
velocity in porous media is lower than a minimum value, it cannot be treated by
Darcy’s law, and a linear relationship between pressure gradient and velocity does
not exist. This is called pre-Darcy flow [9, 14–17]. A simple interpretation of this
concept is that there is no need to expect the linear relationship between velocities
versus pressure gradient to pass through the origin. In other words, care should be
paid when thinking of Darcy equation as general flow mechanism in porous media,
as this equation may apply at low velocity single-phase flow of some fluids in some
cases, but not to other fluids in other porous media. Both pre-Darcy and post-Darcy
3 Basics of Oil and Gas Flow in Reservoirs 79

flowing conditions may be attributed as non-Darcy Flow, although the post-Darcy,


i.e. high flow velocity, has been studied by many researchers. Studies of Prada and
Civan [18] revealed that pressure gradient in porous media should be higher than a
threshold for the Darcy’s law to apply.
The fluid force needs to be sufficient to overcome threshold pressure gradient.
They suggest that Darcy’s law needs correction to account for this effect. Evidences
of pre-Darcy flow were reported by others, mostly in the fields of water and soil
science and engineering [19–23]. This concept is emerging in petroleum reservoir
engineering, and numerical software has been developed to study low-velocity flow
behavior in porous media. For example, Yu et al. [24] applied a low-velocity, non-
Darcy flow numerical simulator to study a low-permeability oil formation. Farmani
et al. investigated the pre-Darcy flow for gas and liquid phases in unconsolidated sand
pack. N-heptane, n-hexane, water and condensate were used as liquid phase and N2 ,
CH4 and CO2 were used as gas phase. They observed pre-Darcy, Darcy and post-
Darcy flow regimes for these substances. Also, they investigated these flow regimes
in a porous media with different particle size. They reported that the pre-Darcy flow
depends on different parameters such as low velocity, porous medium properties,
fluid properties and interaction between medium and fluid. Also, the mobility of the
system was found as an important parameter on the onset of pre-Darcy flow [25].
In another work, Farmani et al. investigated the pre-Darcy flow and Klinkenberg
effect in the calcite, dolomite and chalk-salt core samples. These cores have low
permeability (0.011–18.53 md) and low porosity (12.33–28.21%). The onset of pre-
Darcy flow was investigated in these cores. N2 was used as the flowing gas in core
flood experiments. They reported the onset of pre-Darcy flow in each core. Also, the
pre-Darcy, Darcy and post-Darcy flow regimes were determined using the Reynolds
number analysis and reduced pressure drop analysis [26].
Another form of flow through porous media is defined by Brinkman and is known
as Brinkman equation, shown in Eq. (3.5) [27]. Brinkman flow describes a flow
pattern combined of Darcy and viscous flow [3, 7]. The Brinkman equation considers
extra terms in the form of second derivative of velocity to account for additional
pressure drop compared to laminar Darcy pressure drop. This equation applies in
unconsolidated porous medium with large permeability, where the viscous shearing
stresses acting on a volume element of fluid cannot be neglected [27]. This type of
flow frequently occurs when experiments are designed and conducted in unconsoli-
dated sand pack, glass beads, catalyst pack, and etc. where the flow is neither fully
Darcy nor fully viscous. The Brinkman equation approaches Darcy law for low rock
permeability, and approaches viscous flow when permeability approaches infinity.

∂ 2v ∂ 2v v 1 dp
+ 2− = (3.5)
∂x 2 ∂y k μ dz

However, there is no clear idea about the range of flow rates where the flow regime
changes from one state to another state in porous media. According to de Lemos,
a distinction between laminar, nonlinear, and fully turbulent flow in porous media
is not as evident as it is observed in unobstructed flow, and adequate models which
80 R. Azin et al.

cover a wide range of medium (K, ϕ) and flow properties (ReH ) still have to be
developed [28]. Dybbs and Edwards classified the flow regimes into four regions,
starting with laminar or Darcy flow, which is characterized by Re = 1 to turbulent
flow characterized by Re > 300 [29]. The modified Reynolds number based on Ergun
equation takes into account the porosity and is used to calculate the friction factor
for flow through porous media [29].
As discussed earlier, Darcy’s law arises when the flow on the microscopic scale is
Stokes flow, which is governed by the Navier–Stokes equations without the nonlinear
advective terms. Under these conditions, the microscopic Reynolds is negligible.
Once the microscopic Reynolds number becomes sufficiently large, then steady recir-
culating regions form downstream of particles, and even before this happens, the flow
past, say, a regular array of cylinders or spheres, become non-symmetric due to the
increasing influence of the advective terms in the Navier–Stokes equations. This is
still well within the laminar regime and the Reynolds is still single digit or perhaps
a little larger. Therefore, it seems that while turbulent microscopic flow is definitely
non-Darcy, non-Darcy flow is not necessarily turbulent and there remains quite a
large range of values of Re for which the microscopic flow remains laminar. In other
words, the non-Darcy flow does not necessarily correspond to change in flow regime
from laminar to turbulent flow, and deviation from linear Darcy flow is not initiated
by turbulence. As will be discussed later, the onset of non-Darcy flow occurs at
Reynolds numbers much lower than the onset of turbulent flow.
Also, Hubbert addressed some of the Darcy’s law deficiencies [30]. He performed
a dimensional analysis on Darcy equation and explored that this equation is a purely
kinematic expression in its primitive form which involves explicitly only the dimen-
sions of length and time, while the flow of a viscous fluid through a three-dimensional
network of channels in porous solids is a dynamic phenomenon that involves forces
and energies. According to Hubbert, it is not clear how the fluid properties, density
and dynamic viscosity are involved in the flow, and how the flow rate is affected
by the geometrical properties of the porous solid [30]. There are many forms of
Darcy equation, each referring to a special gradient form. The two common forms
of Darcy equation are expressed as pressure gradient and fluid potential gradients
[5]. Hubbert introduced the potential form of Darcy’s law in terms of energies and
forces, although the potential is indeterminate or nonexistent for some cases. The
potential form of Darcy’s law and the concept of equipotential surfaces are used to
analyze such phenomena as oil migration, hydrocarbon entrapment, identification of
oil–water interface, and etc. [30].
In addition to the pre-Darcy flow, there are other conditions where flow through
porous media, in this case oil reservoirs, may deviate from Darcy’s law. This
phenomenon may occur when flow velocity increases, so that the linear relation-
ship between flow rate and pressure gradient will be no longer valid. Rather, flow
equation is parabolic at high velocities and occurs mostly near wellbore. In the context
of fluid mechanics, this phenomenon may be expressed as transition from laminar
to turbulent flow, as expressed similarly in the context of reservoir engineering [7,
11]. The non-Darcy flow observed in oil and gas reservoirs implies a high pressure
gradient near wellbore, which causes the diffusivity equation to be non-linear at this
3 Basics of Oil and Gas Flow in Reservoirs 81

region [13]. The transition from Darcy, i.e. linear flow, to non-Darcy, i.e. turbulent
flow depends on reservoir rock and fluid properties and oil production rate. It should
be noted that the transition or change in flowing regime from Darcy to non-Darcy
depends mainly on well production rate and reservoir permeability. For oil reservoirs,
Azin et al. showed that non-Darcy flow is expected at flow rates above 500 bbl/day
[13].
In general, there are two criteria which determine flow regime and transition from
Darcy to non-Darcy flow in porous media. These criteria are the Reynolds number and
the Forchheimer number. The Forchheimer number represents the ratio of pressure
drop consumed by liquid–solid interactions to that by viscous resistance. This number
has a direct relation to non-Darcy effect [31]. The critical Forchheimer number is
a criterion for the onset of non-Darcy flow and may be selected according to the
features of the problem, and should be scientifically more reasonable in compared to
a fixed critical value. Zimmerman et al. performed flow experiments and simulations
to evaluate transition between Darcy-non-Darcy flow regimes [32]. They defined a
critical Reynolds number, at which the non-Darcy pressure drop contributes 10%
of the total pressure drop. For practical purposes, this transition seems to occur at
about Re = 10 [32]. Hassanizadeh and Gray conducted a survey of the literature to
reveal that the critical Reynolds number which determines the transition between
Darcy and non-Darcy flow ranges between 1 and 15 [33]. Zeng and Grigg reviewed
the two types of non-Darcy criteria and proposed a revised Forchheimer number
[31]. They suggested a reference value of 0.11 for the critical Forchheimer number,
which corresponds to a 10% non-Darcy effect [32]. In another study, Hassanizadeh
and Gray showed that the onset of nonlinearities occurs at Reynolds numbers around
10, where macroscopic viscous and inertial forces become negligible in comparison
with microscopic viscous forces [33]. They analyzed different theories, empirical
correlations, and mechanisms that support the non-Darcy flow. They suggested that
microscopic viscous forces, i.e. drag forces, grow at high flow velocities, so that the
solid–fluid drag force becomes much larger than inertial and viscous stress forces at
the onset of nonlinear flow.

3.2.2 Boundary Determination Between Darcy, Brinkman


and Viscous Flow

Zahedizadeh et al. solved the main flow equations (Darcy, Brinkman and Viscous) in
cylindrical systems. The solution of Viscous and Brinkman equations are presented
in Appendix 3.1 and 3.2, respectively. In their study, the boundary of each flow
is reported. They expressed that flow regime is affected by variation of radius and
permeability. Other characteristics like pressure drop and viscosity do not have any
impact on the flow regime [34].
Figure 3.1 shows the viscous-Brinkman and Darcy-Brinkman flow regime bound-
aries as a function of permeability and radius for different deviation of resulted
82 R. Azin et al.

1E+11
1E+10 S=0.1 Viscous-Brinkman
1E+09 S=0.5 Darcy-Brinkman
100000000 S=1
10000000 S=2
1000000
100000
10000
K (md)

1000
100
10
1
0.1
0.01
0.001
Darcy Flow
0.0001
0.00001
0.1 1 10 100 1000
Radius (μm)

Fig. 3.1 Darcy-Brinkman and viscous-Brinkman flow regime boundaries in cylindrical coordinate
for different deviation of velocity profiles (S) [34]

velocity profiles (S). Based on this figure, by increasing R at a constant permeability,


the type of flow regime will change from viscous to Brinkman and then Brinkman
to Darcy. The viscous flow is dominated in the small radius due to the effect of wall
on the velocity profile. The wall effect is considerable only on the boundary layer
in large flow path dimensions. In these kinds of flow path, most parts of the velocity
profile are not affected by the wall. So, the Darcy flow will be dominated in these
conditions. Based on Fig. 3.1, by increasing the S parameter in the viscous-Brinkman
flow boundary, the flow type changes to Brinkman flow. Therefore, S = 0.1 indicates
the exact boundary of viscous and Brinkman flow [34].
In Fig. 3.2, the permeabilities related to the different S parameter values for a
constant radius (R = 10 μm) in the Brinkman-viscous flow boundary are obtained
by using Fig. 3.1. Figure 3.2 shows that the Brinkman velocity profile approaches the
viscous flow velocity profile by decreasing the S parameter. At S = 0.1, the velocity
distribution diagram which calculated from the Brinkman and viscous equations
have a high agreement. Therefore, in S ≤ 0.1, the viscous equation can be employed
instead of the Brinkman flow equation [34].
Based on Fig. 3.1, by an increase in the S parameter in the Darcy-Brinkman flow
boundary, the flow regime type tends to Brinkman. In this condition, it can be stated
that S = 0.5 shows the exact boundary of Darcy and Brinkman flow regimes. In
Fig. 3.2, the permeabilities related to the different S values for a constant radius (R =
10 μm), in the Brinkman-Darcy flow regime boundary are gained using Fig. 3.1, and
the velocity profile is achieved, based on the Brinkman equation. Also, the velocity
profile for the Darcy flow regime is shown for permeability values related to different
S values [34].
3 Basics of Oil and Gas Flow in Reservoirs 83

Fig. 3.2 Viscous velocity


profile and Brinkman
velocity profile for different
deviation of velocity profiles
(S) in cylindrical coordinates
[34]

Based on Fig. 3.3, it can be seen that the effect of the boundary layer reduces
by decreasing the S parameter. In this case, the Brinkman velocity profile tends to
Darcy velocity. At S = 0.5, the fluid flow velocity profile achieved from the Brinkman
equation corresponds to the Darcy equation in more than 99% of the dimensionless
radius. Therefore, the wall effect is at its minimum value in S ≤ 0.5 values, and the
flow regime can be assumed as Darcy flow. In this condition, in the regions, lower of
the line S = 0.5 in the Brinkman-Darcy flow boundary in Fig. 3.1, the Darcy equation
can be used to calculate the fluid flow [34].

Fig. 3.3 Darcy and 4


Brinkman velocity profile for S=2
different S values in 3.5
cylindrical coordinates [34]
3 Darcy flow
Brinkman flow
Velocity (cm/hr)

2.5

2 0.1
1.5 R=10 μm
1 S=1

0.5
S=0.5
0
0.9 0.95 1
r/R
84 R. Azin et al.

3.2.3 Flowing Fluids and Diffusivity Equation in Porous


Media

The flowing fluid in reservoir may be incompressible, slightly compressible, or


compressible in nature. The oil flowing through porous media is regarded as slightly
compressible, and its density changes slightly with pressure. This is especially impor-
tant for flow of oil in reservoir, as the reservoir is typically subject to pressure change,
normally pressure depletion. Therefore, a slightly compressible fluid tends to expand
upon pressure depletion and occupy more space at lower pressure. Expansion of
fluid will assist in oil recovery, as the reservoir volume is assumed constant, so the
expanded fluid will contribute to reservoir drive and pressurize the formation. Reser-
voir engineers apply a simple exponential equation to define pressure dependency
of reservoir fluid [35]. This equation is applicable at constant temperature, typically
for thin formations where temperature gradient is negligible. In addition, the oil flow
may occur as single phase or associated with one (water/gas), two (water and gas), or
even more (water, supercritical fluid, gas, solids like precipitated but not deposited
asphaltene and scale) phases. The two-phase and three-phase flow through porous
media has been more addressed in the literature. It should be mentioned that the oil
flow may be originally single phase, but switch to two-phase as a result of pressure
depletion and release of solution gas. In this case, single phase flow continues until
the gas saturation rises beyond critical gas saturation and gas starts flowing with oil.
The flowing gas may have been injected for pressure maintenance or EOR purposes.
In addition, water can flow with oil as separate phase. Sources of water are injected
water for displacement (EOR) and/or pressure maintenance, and natural water influx
from aquifer. There are cases when three phases are present and flow all together,
e.g. in the case of combined water and gas injection in water alternating gas (WAG)
process, or in the case of natural water influx into a saturated oil reservoir. Generally,
analysis of fluid flow becomes more complex when the number of flowing phases
increase. For example, density difference between flowing phases may cause gravity
segregation, lead to gas or water coning, improper oil sweep, and trapping problems.
Geometrically, flow may occur in linear, radial, spherical or semispherical forms.
Radial flow is a common flow geometry, as it resembles oil flow towards a producing
well. In the case of fluid flow through naturally or induced fractures, linear flow
is dominant. Also, when a well is partially penetrated into producing formation
or partially perforated, spherical flow becomes dominant. Many classic reservoir
engineering textbooks describe different flow geometries, of which the works of
Craft et al. [35], Dake [5], and Ahmed [8].
The fluid velocity in reservoir depends on the type of fluid, its viscosity and produc-
tion rate at the well. For radial configuration, the fluid velocity is low at distances far
from well, and increases as fluid approaches wellbore. Generally, flowing velocity is
higher in gas reservoirs compared to oil reservoirs. Longmuir calculated superficial
velocity for a number of oil and gas wells [9]. Typical fluid velocities at a radius
of 10 ft from wellbore were about 275 ft/day, 7 ft/day, and 0.1 ft/day for gas well,
high-rate oil well, and low-rate oil well, respectively. The fluid velocities decreased
3 Basics of Oil and Gas Flow in Reservoirs 85

drastically at distances far from wellbore. For the same reservoirs, velocities at a
radius of 700 ft from wellbore were calculated about 3 ft/day, 0.11ft/day, and 0.002
ft/day for gas well, high-rate oil well, and low-rate oil well, respectively.
Production of oil from reservoir depends basically on flow of oil through porous
reservoir. The wells drilled in the formation need be fed by fluid to produce hydro-
carbon. Flow of fluid through porous medium is the fluid pass towards well. Under-
standing fundamentals of fluid flow and flow mechanisms are essential for proper
reservoir management. Flow of fluids through porous media is associated with
changes in fluid pressure as a function of time and position, in the case radial flow as
a function of time and radius with respect to production well. This pressure behavior
in oil reservoir is described in the form of diffusivity equation derived from material
balance or continuity equation in the reservoir. The linear form of diffusivity equation
is shown as Eq. (3.6):

∂2 p 1 ∂ p φμct ∂p
+ = (3.6)
∂r 2 r ∂r 0.0002637k ∂t
Equation (3.6) is written in radial form, as is the usual flow pattern in oil flow
towards a producing well. The linear diffusivity equation is the simplified form of
general diffusivity equation in which the non-linear term is removed for simplicity.
This equation is composed of continuity equation combined with Darcy’s law as the
auxiliary equation. The diffusivity equation shown in Eq. (3.6) is written for slightly
compressible fluid. Equation (3.6) is linear, which implies that pressure distribution
during oil flow can be described by a linear PDE. This equation can be solved for
different boundary conditions that define transient and pseudo steady state solutions.
This is why the diffusivity equation is regarded as the basis for almost all reservoir
engineering calculations, including pressure distribution in transient and pseudo-
steady state conditions, well test analysis, and etc. Analytical solution of the linear
diffusivity equation is given in Appendix 3.3.
The nonlinear PDE form of diffusivity equation arises when pressure gradient, ∂∂rp ,
in reservoir becomes significant. In terms of velocity, Darcy velocity is determined
directly by pressure gradient. A higher ∂∂rp implies larger Darcy velocity. Meanwhile,
a larger flow velocity which is frequently the case in near-wellbore region, implies
nonlinear PDE as shown in Eq. (3.7):
 
∂2 p 1 ∂ p 1 ∂ρ ∂ p 2 φμct ∂p
+ + = (3.7)
∂r 2 r ∂r ρ ∂ p ∂r 0.0002637k ∂t

It is worth noting that a nonlinear diffusivity equation does not mean that there
is a change in flow regime from laminar to turbulent; rather, it just implies a higher
pressure gradient and flowing velocity which is still within the laminar flow regime.
The analytical solution of Eq. (3.7) is presented in Appendix 3.4.
86 R. Azin et al.

3.2.4 Relative Permeability

Although the Darcy equation was developed for single phase flow, it was later
extended to formulate and model multiphase flow through porous media by intro-
ducing the effective and relative permeability concepts. Multiphase flow in porous
media is significant in Soil Science, Hydrology, Chemical Engineering and Petroleum
Engineering applications. For reservoir engineers, multiphase flow is essential in
reservoir simulation, determining reservoir performance, well productivity, well
injectivity and reservoir ultimate recovery. Phenomena such as diffusion, dispersion,
viscous fingering are more dominant in porous medium especially in the presence of
heterogeneities. These factors are the main reasons for complexity of multiphase flow
in porous media. When there is more than one phase flowing through porous media,
each phase flows by its effective permeability, and saturation of one phase affects
rock permeability to other phases. Effective permeability is a relative measure of the
conductance of the porous medium for one fluid while it is saturated with more than
one fluid [8]. Thus, for example, the porous medium has an effective permeability to
oil and an effective permeability to water in the two-phase oil–water flow. The ratio
of effective to absolute permeability is called relative permeability. This parameter
is a function of fluid saturation, rock and fluid properties, including pore structure,
pore size distribution, porosity, absolute permeability, saturation history, wettability
of surfaces, interfacial tension, viscosity of fluid, temperature, flow rate, and over-
burden pressure [36–38]. Relative permeability data are shown either in tabulated
form or as relative permeability curves which are plotted against fluid (usually the
wetting phase) saturation. For oil/water systems, relative permeability curves are
tabulated and plotted against water saturation, for gas/oil and gas/water systems,
these data are demonstrated against liquid saturation. The relative permeability data
are generated either in laboratories using the analysis of multiphase fluid flow in the
core or predicted through mathematical correlations and models. Overall, labora-
tory methods are categorized into two major groups, steady state and unsteady state
methods [39]. In steady state method, the immiscible fluids flow simultaneously in
core plugs until saturation and pressure equilibrium is attained. The unsteady state
method involves injecting a fluid at constant rate or constant pressure to displace in-
situ fluid. Laboratory measurements of relative permeability curves are usually very
sensitive, time consuming and costly [40]. Hence, researchers prefer to obtain these
data from other methods that are quick and accurate. Dake discussed the challenges
that arise during laboratory measurements of relative permeability, mainly due to
differences between mobility ratio of flowing phases and reservoir heterogeneity
[41].
In addition to laboratory measurements, empirical and analytical mathematical
correlations have been proposed yet for predicting or estimating the relative perme-
ability of two-phase systems, that each of them has its own characteristics. Empirical
correlations and analytical mathematical models are widely used for predicting rela-
tive permeability data. Generally, most of these models have been developed as a
3 Basics of Oil and Gas Flow in Reservoirs 87

function of wetting phase saturation by using a stepwise linear and nonlinear regres-
sion analyses or analytical/numerical approach such as capillary models, statistical
models and network models which are based on the assumption that a porous media
consists of a bundle of capillary tubes. Mohammadi-Baghmolaei et al. reviewed
a number of common approaches for relative permeability prediction models and
correlations. They stated that the relations for estimation of relative permeability of
two phase systems are divided into two categories. The first category includes such
correlations which predict relative permeability as function of capillary pressure,
absolute permeability, porosity, fluid saturation, interfacial tension and etc. [42]. The
following correlations are in the first category which summarized in Table 3.1.
Purcell in 1949 introduced the first analytical mathematical models to estimate the
relative permeability of water–oil and gas-oil systems using capillary pressure data
[43]. In 1951, Fatt and Dykstra presented correlations to calculate the wetting phase
relative permeability using developed equations by Purcell as a base model [44]. They
considered a lithology factor for including a fluid path that is a function of saturation.
In 1953, Burdine developed similar equations with Purcell’s models through the
introduction of a tortuosity factor as a function of wetting phase saturation [45]. In
1954, Corey developed an empirical correlation to estimate relative permeability of
gas-oil systems based on relative permeability measurements on a large number of
cores from several formations [46]. Corey also realized that the capillary pressure
curve could be approximated by a linear equation and replaced in Purcell’s model. In
1958, Wiley and Gardner presented equations to estimate the relative permeability

Table 3.1 The first category of relative permeability correlations


Year Relative permeability System Rock type Wettability References
correlations
1949 Purcell OW/GL – – [43]
1951 Fatt and Dykstra OW Sandstone – [44]
1953 Burdine OW/GL Sandstone – [45]
1954 Corey GL Sandstone – [46]
1958 Wyllie and Gardner OW/GL Sandstone – [47]
1958 Wahl GL Sandstone – [48]
1958 Torcaso and Wyllie GL Sandstone – [49]
1966 Brooks-Corey OW/GL Sandstone – [50]
1982 Honarpour et al. OW/GL Sandstone/Carbonate Any wettability [51]
2000 Ibrahim and OW/GL Sandstone/Carbonate Any wettability [52]
Koederitz
2001 Mulyadi, Amin, and GL Sandstone – [53]
Kennaird
2013 Mosavat et al. OW Sandstone – [36]
2015 Xu et al. OW Sandstone Water-wet [40]
O Oil; W Water; G Gas; L Liquid
88 R. Azin et al.

oil–water and gas-oil systems in drainage process using capillary pressure data [47].
Also, Wahl et al. presented equation to calculate the ratio of relative permeabilities
in gas-oil systems based on sandstone field measurements in drainage process [48].
Torcaso and Wyllie developed a simple formula to calculate the relative permeability
of oil phase in gas-oil systems [49]. Prison presented correlations for estimation
of the relative permeabilities of wetting and non-wetting phases in drainage and
imbibition processes and in clean, water-wet rocks [8]. In 1966, Brooks and Corey
improved Corey’s capillary pressure model and proposed correlations for capillary
pressure and relative permeability of wetting and non-wetting phases [50]. In 1982,
Honarpour et al. developed empirical correlations for water–oil and gas-oil systems
in which the impacts of wettability and rock type had been considered [51]. In 2000,
Ibrahim and Koederitz presented correlations for estimating relative permeability in
4 different systems through implementation of a linear regression model by using
416 sets of measured relative permeability data [52]. In 2001, Mulyadi, Amin, and
Kennaird proposed MAK correlation to estimate relative permeability in the gas–
water system. This correlation is based on Corey’s model for (gas-oil system) [53].
In 2005, Lomeland et al. proposed smooth and flexible three-parameter analytical
correlations for relative permeability of water–oil, gas-oil and gas–water systems.
Their model has flexibility for entire saturation range and also shows good agreement
with experimental data in high or low fluid saturation [54]. The impact of pressure,
fluid viscosity and flow rate were studied by Mosavat et al. [36]. Their proposed
correlation estimates relative permeability of water–oil systems based on Corey’s
model. Xu et al. proposed an empirical model to consider the impact of displacement
pressure gradient for relative permeability of water phase in water–oil system [40].
The second category refers to the parametric models. The required parameters
are obtained by implementing measured experimental data. Indeed, these models
cannot be used in the absence of laboratory measured data. Also, some models have
adjustable parameters that need be tuned based on measured data. As an example,
correlations developed by Corey’s [46], Brooks and Corey [50], Lomeland et al.
(LET) [54], Sigmund and McCaffery [55], Chierici [56, 57], Van Genuchten and
Maulem [58], and Li et al. [59] have one or more adjustable parameters that need be
adjusted before they can be applied to a reservoir model. The information of these
correlations are summarized in Table 3.2.
The decision tree shown in Fig. 3.4 can act as a guideline to select a proper model
to predict relative permeability for a given system. Based on this figure, the LET
model applies in most cases of oil and gas phases, both in carbonate and in sandstone
systems.
In addition to empirical correlations, intelligent systems such as artificial neural
network (ANN), adaptive neuro-fuzzy inference system (ANFIS) and least square
support vector machine (LSSVM) have been applied to predict of petrophysical
properties, including relative permeability. Also, Ahmadi et al. used ANN intelli-
gent model to predict the oil/water relative permeabilities with different learning
algorithms [60]. The gas/oil relative permeabilities were also modelled by Ahmadi
using LSSVM model [61]. Despite this growing rate of predicting petrophysical
properties using intelligent systems, it should be mentioned that dealing with these
3 Basics of Oil and Gas Flow in Reservoirs 89

Table 3.2 The second category of relative permeability correlations


Year Relative permeability correlations System Rock type References
1954 Corey generalized correlation OW/GL – [46]
1966 Brooks and Corey generalized correlation OW/GL – [50]
1979 Sigmund and McCaffery OW/GL Carbonate [55]
1984 Chierici OW/GL Sandstone [57]
1999 Van Genuchten-Maulem OW/GL – [58]
2005 LET OW/GL Composite rock [54]
2006 Li et al. OW – [59]
O Oil; W Water; G Gas; L Liquid

Fig. 3.4 Decision tree for proper model selection of relative permeability

systems cause some severity. The intelligent systems are highly structural-dependent
and require an available source of data for development of models. Moreover,
these systems do not provide an applicable clear mathematical function that can
be employed in modelling and simulation as well. It is worth mentioning that among
intelligent models, those which provide mathematical relationship between input data
and output are more attractive for engineers. This will eliminate the sense of “pas-
sive” role in “black-box” calculations of intelligent techniques and let the user play
a more “active” role and handle equations and correlations in engineering problems.
Mohamadi-Baghmolaei et al. applied the multi-gene genetic programming (MGGP)
approach to develop a new consolidated robust equation which predicts relative
permeability of gas/oil systems [42]. Their proposed correlation requires rock and
fluid properties such as gas saturation, gas molecular weight, oil API gravity, rock
porosity, and absolute permeability. These properties are routinely measured on rock
90 R. Azin et al.

and fluid samples. In the following, the relative permeability correlations obtained by
genetic algorithm are proposed by Farmani et al. and Mohamadi-Baghmolaei et al.
Mohamadi-Baghmolaei et al. gas phase relative permeability correlation [42]:

krg = 0.2308ag − 1.091 (3.8)

2
ag = 4.947( Sg∗ + bg + cg + dg + eg (3.9)
 
0.000993 × f g × A P I + (Ln(A P I ))2
bg = − (3.10)
M Wg
  
 ∗
  kabs
cg = Ln A P I − 5.741Sg Ln M Wg + 0.003574 Ln(Ln(A P I ))
ϕ
(3.11)
4  
0.005159 (M W g 29.5531 + 0.034718 kabs ϕ
+ M Wg
dg = (3.12)
(A P I )4
⎛ ⎞ 41
109.4206 ⎠
eg = ⎝   (3.13)
M Wg kabsϕ
  
f g = A P I E x p −6.996025Sg∗
  4   
+ 9.701419 Sg∗ Ln M Wg Ln(A P I ) Ln(Ln(A P I )) − A P I (3.14)

Mohamadi-Baghmolaei et al. oil phase relative permeability correlation [42]:

kr∗o = 0.09595ao + 0.03296 (3.15)


 
( So∗ +ε)((So +ε) )
∗ co

ao = So∗ (3.057 − bo ) (3.333 − do ) (3.16)

 So∗ 7.841M W
(M W g− M Wg + M Wg −2 A Pg I
bo =   (3.17)
5.504M Wg − 0.7117 kabs
ϕ
+ 2.135A P I + 2.372

7.815 − 2 A P I + eo
co =   (3.18)
4.682M Wg − 0.5836 kabs
ϕ
+ 1.751A P I + 2.006
 
15.47 + E x p 6.143So∗ + f o
do =      
kabs
ϕ
− A P I 0.7117 kabs
ϕ
− 5.504M Wg − 2.135A P I − 5.562
3 Basics of Oil and Gas Flow in Reservoirs 91

Farmani et al. gas phase relative permeability correlation [62]:



krg = 2.284ag + 3.331 (3.19)

 
ag = (20.53 Sg∗0.25 + bg cg dg (3.20)

   μg
bg = [ Sg∗ + 4.105 3.159Sg∗ − 4.159Nca 17.6 − 20.53 − 217.1Sg∗
μo

 ∗
 Sg
+ 3I F T − 4Sg + 4.457 4.159
I F T (k/ϕ − 4.866)(k/ϕ Nca )
 
+ 20.53 k/ϕ)1/8 − 481.8 (3.21)

I FT  
cg = [55.18Sg∗ − 325.1Nca + √ (k/ϕ Nca )
Sg∗ 0.25 − Sg∗ + 4.875
k/ϕ
+ 8.403))4.247 + 21.96(k/ϕ)1/8 − 331.3] (3.22)
⎡    ⎤
 ∗ I F T − μg + 4.402 + 4.159 
μ  μ S g μ 
dg = I F T N ca ⎣ g 
− I F T +
g
+
o
+ 10.53⎦ (3.23)
μo  μo Mwg 

Farmani et al. oil phase relative permeability correlation [62]:

kr∗o = 0.1588ag + 0.07801 (3.24)


 
1.238Nca Nca Nca
ao = ln ( − + 0.1472) do
bo co
   0.221
μg
I F T −5.774 μo −4.889
   
+ exp(Sg∗ ) Nca Sg∗ − 0.221 + exp exp Sg∗
  
+10.59Mwg Nca exp − exp Sg∗ (3.25)
     ∗ 
    Mw  Nca Sg
Sg∗ + 0.4007 Sg∗ + 0.007532
g
bo = + ∗ + (3.26)
kϕ Sg + 0.1472 k/ϕ
    ∗ 
  4.956  
N Sg
co = Sg∗ + 0.1732
ca
− 0.334 (17.98Nca ) + + (3.27)
k/ϕ Sg∗ + 0.1472 k/ϕ

Their results show that these correlation can predict the gas and oil relative perme-
ability with high accuracy. More information on the procedure of proposing these
correlations are available on [42, 62].
92 R. Azin et al.

3.2.5 Flow Through Fractures and Fissures

Natural fractures play crucial role in facilitating flow in tight reservoirs. Depending
on the size and extension of fractures, analysis of fluid flow through these conduits
may be treated as viscous flow (Navier–Stokes approach), Darcy flow, or Brinkman
flow. When the size of fracture approaches typical pore size, fluid flow can be treated
as it was Darcy in porous media. On the other side, the flow of fluid will be treated
as viscous flow when the fracture opening is macroscopic with minimized capillary
effect. The Brinkman flow occurs in fracture opening between porous and macro-
scopic scales. This type of flow is frequently observed when the researchers conduct
flow experiments on unconsolidated sand pack, and some may wrongly interpret the
results simply by Darcy law. The viscous flow gradually converts to Brinkman and
then to Darcy flow as the fracture opening decreases, and for fractures opening at the
order of one micrometer, the flow is almost completely Darcy type [3, 7]. A common
mistake to reservoir engineers is that they assume Darcy flow in fractures, which
arises when they equate the flow equation derived for a single fracture to Darcy flow
in order to obtain fracture permeability. This approach implies that the flow through
fractures is always of Darcy type, regardless of fluid type and fracture dimensions.
Moreover, the apparent permeability obtained by this approach is valid for a single
fracture and is not applicable for a set of fractures or a fracture network. Golf-Racht
proposed a different equation for average permeability of a fracture set, which is
typically different from that of a single fracture [63].
Barenblatt et al. introduced the concept of dual porosity. Based on this concept,
the fracture is the main flow path of the fluid, however, the matrix stores the hydro-
carbon [64]. Warren and Root modified the Barenblatt model by using the fracture
compressibility. In this model, fluid storage capacity (ω) and inter-porosity flow (λ)
[65]. Zahedizadeh et al. provide a solution for Warren and Root model based on the
reservoir with different boundary conditions [66].
A schematic of fractured reservoir and its boundaries is shown in Fig. 3.5. The
Warren-Root model obtains pressure distribution in fractured reservoirs Eqs. 3.28
and 3.29 introduce the fluid flow mathematical model through matrix and fractures
[63]:

 m = − Km grad(Pm )
U (3.28)
μ

 f = − Kf grad(Pf )
U (3.29)
μ

Fractures have a great impact on the porous medium permeability. Great knowl-
edge of fracture network and flow equations are required for the reservoir simulation.
There are two main methods to model the fluid flow through a fractured reservoir
known as equivalent continuum modelling and discrete fracture modelling (DFM)
[67]. Warren and Root introduced the dual porosity concept as one of the basic
equivalent continuum modelling methods to simulate the fractured reservoirs. In the
3 Basics of Oil and Gas Flow in Reservoirs 93

Fig. 3.5 Schematic of the


fractured reservoir model

Warren-Root model, the flow in both fracture and matrix is assumed as Darcy flow
[65]. However, the fluid flow in the fractures is more complex due to the high velocity.
In the later works, a non-Darcian flow was considered as a modification in the calcu-
lations of the dual porosity/dual permeability. Choi et al. employed the Forchheimer
equation as a non-Darcy term in the fractures. This term shows that the results
of these two models differ significantly. However, the results of these two models
show good agreement in some specific range of parameters [68]. DFM is known
as one of the most suitable methods for explaining the explicitly of fracture and
matrix systems [69, 70]. Nevertheless, the main disadvantage of this method is that
meshing a complex three-dimensional fracture network is very difficult. Also, this
method has high computational cost due to a huge number of grids and convergence
problems. Therefore, DFM application in the field scale is not realistic at present.
But it seems that the DFM is becoming more applicable in the future according to
its improvements [71, 72].
Equation (3.30) is resulted as a reservoir dimensionless pressure equation in
Laplace space:

PDf (z, rD ) = C1 I0 (M rD ) + C2 K0 (M rD ) (3.30)

where, C 1 and C 2 are defined as the Bessel equation constants. These constants are
calculated by applying boundary conditions. I 0 and K 0 √ are first and second type
modified Bessel functions in zero order. M is defined as zf(z). This parameter is
used to simplify the calculations. F(z) is shown mathematically by Eq. (3.31):

ω(1 − ω)z + λ
f(z) = (3.31)
(1 − ω)z + λ
94 R. Azin et al.

Table 3.3 demonstrates the solution of Eq. (3.30) by different inner and outer
boundary conditions. The solution of Eqs. 3.32 through 3.37 will be presented in
Appendix 3.5.
Figure 3.6 indicates the result of production behavior in various outer and inner
boundary conditions. The first region in Fig. 3.6 shows the production at initial times,
and production of fluid is mainly through fractures. The second region in Fig. 3.6a
shows the transition zone. In this region, both fractures and matrix contribute to
fluid production. Also, the dimensionless pressure (PD ) stays almost constant in
this region. In the transition zone and in the case of constant pressure production
(Fig. 3.6b), both fractures and matrixes play the same role in fluid production where
the dimensionless rate (qD ) stays almost constant. In the third region, the matrix has a
higher contribution in reservoir production rate because of fracture capacity for fluid
storage ends up. The fourth region in Fig. 3.6 refers to reservoir boundary effect, and
its behavior depends on outer boundary condition.
Figure 3.7a, b show the effect of dimensionless reservoir radius (rDe ) on wellbore
pressure profile and production rate, respectively. This figure demonstrates that the
onset of region IV depends on the selected reservoir radius. The reservoir boundary
effect gets away from the transition zone, as the reservoir radius increases.
Figure 3.8 shows the plots of PD -tD and qD -tD at constant rDe and ω for closed
outer boundary condition. According to these figures, by reducing the matrix-fracture
inter-porosity flow (λ), the transition zone accosts the reservoir boundary effect. High
λ values imply high matrix ability to move fluid toward the fractures and show higher
matrix to fracture permeability ratio. Therefore, a lower pressure difference is needed
to remove fluid from the matrix to fracture in high λ values. Figure 3.8 also indicates

Table 3.3 Solution of PDE for fractured reservoir with different inner and outer boundary
conditions
Inner BC Outer BC Solution Equations
Constant production Closed PDf (z · rD ) = (3.32)
rate K1 (M rDe ) I0 (M rD )+I1 (M rDe )K0 (M rD )
z M [K1 (M) I1 (M rDe )− K1 (M rDe ) I1 (M)]

Constant production Constant PDf (z · rD ) = (3.33)


rate pressure K0 (M rD )I0 (M rDe )−K0 (M rDe )I0 (M rD )
z M[K0 (M rDe )I1 (M )+K1 (M )I0 (M rDe )]

Constant pressure Closed qD (z · rD ) = (3.34)


production M[ K1 (M) I1 (M rDe )− K1 (M rDe ) I1 (M)]
z[K1 (M rDe ) I0 (M rD )+I1 (M rDe )K0 (M rD )]

Constant pressure Constant qD (z · rD ) = (3.35)


production pressure M[K0 (M rDe )I1 (M )+K1 (M )I0 (M rDe )]
z [K0 (M rD )I0 (M rDe )−K0 (M rDe )I0 (M rD )]

Constant production Infinite Laplace space form: (3.36)


rate K0 (M rD )
PD (z · rD ) = z M K1 (M)
Warren-Root analytical solution: (3.37)
PD (tD · 1) = 21 Ln (2.25tD ), M <
0.01
3 Basics of Oil and Gas Flow in Reservoirs 95

Fig. 3.6 Different regions in a—constant production rate, b—constant pressure production

Fig. 3.7 Different reservoir dimensionless radius in a—constant production rate, b—constant
pressure production

Fig. 3.8 Different λ values in constant rD and ω for closed outer boundary condition a—constant
production rate, b—constant pressure production
96 R. Azin et al.

Fig. 3.9 Different ω values in a—constant production rate, b—constant pressure production
condition

that the transition zone interferes with reservoir boundary effect in λ smaller than
10–9 . Therefore, it may be hard to specify the transition zone in the PD -tD or qD -tD
curves in a finite natural fractured reservoir with low λ values.
Figure 3.9 shows the PD -tD and qD -tD curves for different ω values at constant
production rate and constant pressure production conditions, respectively. The tran-
sition zone extends by decreasing fracture storage capacity (ω), as shown in Fig. 3.9.
Based on this figure, the reservoir boundary effect and transition zone do not interfere
with variation of ω.
As mentioned, only λ variation can influence the transition zone and the reservoir
boundary interference. It is impossible to obtain the transition zone and fractures,
when interference occurs between the reservoir boundary effect and transition zone
in a finite reservoir. Prediction of the interference area between reservoir boundary
and transition zone is discussed by Zahedizadeh et al. They introduced λmin for this
purpose where, at λ values less than λmin , it is impossible to identify fractured finite
reservoirs. In other words, λmin for a certain rDe is the smallest λ value in which the
fractured reservoirs are identifiable.
Figure 3.10 indicates the λmin as a function of rDe in a logarithmic scale which
shows the interface points of the reservoir boundary effect and transition zone.
The relationship between λmin and rDe can be written as Eqs. (3.38) and (3.39) for
constant production rate and constant pressure production conditions, respectively:

30.8319
λmin = 2.0599
(3.38)
r De
12.4394
λmin = 2.0001
(3.39)
r De

In the regions above the curve, i.e., λ > λmin , fractures can be detected. On the
other side, in the regions below the curve, i.e., λ < λmin , the reservoir fractures cannot
be detected as usual.
3 Basics of Oil and Gas Flow in Reservoirs 97

Fig. 3.10 λmin values in different rDe in a—constant production rate, b—constant pressure
production condition

3.2.6 Forces Acting on Fluid Flow in Oil Reservoirs

Flow of fluids is governed by forces acting on fluid in reservoir. As expressed before,


the main forces acting on a flowing continuum in classical fluid mechanics include
convection, molecular transport, and external forces [1]. In the case of fluid flow in
reservoir, gravity, capillary, and buoyant forces are active as well.
There are three main forces acting in an oil reservoir, i.e. viscous forces, capillary
forces, and gravity forces. Each force may originate from certain conditions. For
instance, capillary forces arise when two or more phase are present in the capillary
pore space, and may support or resist flow of each phase, depending on the wetting
tendency of rock and type of flowing fluid. The capillary forces and capillary pressure
is frequently defined by Young–Laplace equation:

2σ cos θ
pc = (3.40)
rc

According to this equation, Pc is a function of interfacial tension, σ, rock wetta-


bility, θ, and pore radius, rc , which is a measure of pore size, pore size distribution, and
permeability of porous media. Capillary forces are believed to be a main mechanism
in oil migration and formation of oil reservoirs. In tight reservoirs with low pore radii
and low permeability, capillary effects and capillary forces become more dominant.
In addition, capillary forces provide capillary continuity in oil flow through frac-
tured reservoirs [3], which is also known as block-to-block interaction and matrix
reinfiltration. As a result of capillary continuity, the oil flowing out of a matrix block
re-enters another matrix block rather than flowing through a fracture network towards
producing well. This is mainly due to wetting tendency of matrix block and preferred
wettability towards oil.
Capillary forces are observed in the pore scale, while such phenomena as Buoy-
ancy forces originate from geothermal and pressure gradients in reservoir, which
cause changes in fluid density and result in convective flow in reservoir, provided
98 R. Azin et al.

there is sufficient fracture system available in reservoir. According to Saidi [3],


thermal convection is not likely to occur in an oil field lacking fracture system, even
though it has quite large matrix permeability as high as 1 Darcy.
Gravity flow is an important production mechanism especially in fractured reser-
voirs with low matrix permeability. Oil production rate under gravity flow depends
on density difference between oil and gas, oil viscosity, elevation between the two
contacts, rock permeability, and capillary holdup. The driving force for gravity
drainage is density difference between phases, which provides a pressure difference
and causes instability within the oil column. Although it is theoretically possible that
such an instability occurs in any porous medium, gravity flow is practically observed
in fractured formations only, and presence of fractures is crucial in assisting oil
recovery under gravity drainage [3].
In a conventional reservoir the capillary pressure curve controls fluid distribution
through the reservoir; therefore, the transition zone will come between the water–
oil contact and the oil zone which, in the case of important capillary forces (tight
formation), may be very thick. In a fractured reservoir this situation is completely
different. The discontinuity of the matrix caused by the fracture network cutting the
continuum of the matrix bulk into small individual matrix blocks, explains why the
water table is only related to the fracture network.
In addition, since the fractures are large channels with negligible capillary forces,
the transition zone disappears in a fractured reservoir, and water–oil contact becomes
a horizontal plane. On the other hand, capillary and gravitational forces (through
the capillary pressure curve and gravitational curve) control the static and dynamic
equilibrium of each matrix block. The basic element which relates individual block
behavior to reservoir behavior is the water–oil contact in fractures and is called
water table level. These water–oil contacts in fractures, together with the oil–water
contacts inside the matrix, the last corresponding to displacement front level, are
essential reference planes for the evaluation of the driving mechanism of capillary
and gravity forces. An analogical situation will take place in the case of a gas-cap for
both gas-oil contacts in fractures and matrix blocks, where the first is called gas-cap
table and the second gas displacement front.
If there are more than one force dominating fluid flow, their relative magnitude
would be important to reservoir engineer. For example, when more than one phase
flows through porous medium, viscous and capillary forces are active. In this case,
the ratio of viscous to capillary force makes a dimensionless number called capillary
number, which determines the relative magnitudes of these forces in reservoir. Bond
number is another common dimensionless number, defined as the ratio of gravity
to capillary forces. These are frequently used to interpret EOR experiments and
water flooding operation, although there are questions on adequacy and theoretical
validity of them [3]. For instance, increasing capillary number, i.e. viscous forces
against capillary forces, has been well known as a strategy for increasing ultimate
oil recovery and recovery factor during an EOR process. Determining a threshold
value for each dimensionless number for a certain process depends on rock, i.e.
porosity, permeability, wetting tendency and fluid properties, as well as experimental
conditions. In one experiment, the critical capillary number required to remobilize the
3 Basics of Oil and Gas Flow in Reservoirs 99

trapped phase was determined to be 2E−8 in gas–water systems. A similar order was
observed to remobilize trapped gas in gas-oil systems [73]. Løvoll et al. conducted
experiments on unconsolidated glass beads to investigate the competition of gravity,
capillary and viscous forces during drainage [4]. The synthetic porous medium had
porosity of 63% and permeability of 1915 D, quite high compared to that for normal
oil reservoirs. Although they described some scaling equations and proposed stability
criteria based on generalized Bond number, they did not provide a clear distinction
between gravity, capillary and viscous forces in their experiments.
Convection arises as two modes, Benard convection and Taylor convection [3].
Benard convection takes place when a liquid filled in a container is under negative
density gradient due to either thermal or diffusion process. According to Saidi [3],
Benard convection starts from a stable to unstable conditions, while Taylor convec-
tion starts from an unstable and goes to stable conditions. Convective flux may
occur as a result of concentration gradient, pressure gradient, temperature gradient,
or a combination of two or more gradients. Existence of one gradient can make
another gradient(s) to emerge. For example, combination of pressure and temper-
ature gradient in an oil reservoir may result in compositional changes throughout
the whole depth of reservoir, which is normally known as compositional grading or
compositional gradient [74]. Kiani et al. studied the two dimensional compositional
gradient in mixtures of oil and gas. They also improved a thermodynamic model for
prediction of the methane and plus fraction distribution and compositional gradient
in reservoirs. They used the data of a giant gas condensate reservoir to validate
their model. Their results showed that the developed model could accurately predict
the experimental data. Also, they reported that natural convection could change the
vertical and horizontal compositional gradient and also it affects the pattern of species
distribution [75].
Another example of convection arises as a result of density differences caused
by concentration gradient during CO2 disposal into saline aquifer [76, 77]. Whatso-
ever the origin of gradient, convection is described by Rayleigh number, Ra, which
describes the ratio of convective to viscous forces. Saidi demonstrated convective
flow and resulting compositional gradient in a number of Middle East oil fields
and calculated Ra in these fields [3]. The minimum Ra, known as critical Ra,
required for onset of convection depends on the flow geometry and boundary condi-
tions of problem under study. The convective flow may couple with other trans-
port phenomena, i.e. mass transfer and/or heat transfer. This coupling is known as
thermal diffusion, cross effect, Soret effect or Dufour effect [1]. Thermal diffusion
ratio, thermal diffusion factor, and Soret coefficient are used to determine molecular
flux under the coupled influence of heat and mass transfer [1]. The pressure gradient
can have a more complex impact on oil reservoirs, as the amount of solution gas
dissolved in oil is directly dependent on pressure and temperature. It is well estab-
lished that convective flux acts as a driving force for mass transfer in oil at different
points of reservoir, which in turn imposes compositional gradient characterized by
variations in fluid bubble point pressure, phase diagram, composition, density, and
other physical properties. Ra is sometimes expressed as the product of Grashof and
Schmidt numbers [1]:
100 R. Azin et al.

Ra = Grw · Sc (3.41)

Grw is the diffusional Grashof number, arises due to buoyant forces caused by
concentration inhomogeneities. Sc is defined as the ratio of viscous to diffusive
forces, also known as the ratio of momentum diffusivity to mass diffusivity [1].
Molecular diffusion is normally neglected in oil production and its magnitude is
small compared to other forces. In many practical field operation, impact of this
parameter is considered as minimal due to relatively high production and/or injection
rates in field scale. However, it can compete with other mechanisms in certain cases.
Significance of diffusion in oil recovery from oil reservoirs has been addressed in
the literature, e.g. [3, 7, 78, 79]. Molecular diffusion is also important when applying
solvent-based heavy oil recovery [80–82].
Peclet Number, Pe, is another important dimensionless number which compares
the relative magnitude of convective to diffusive forces in porous medium. Pe arises
when molecular diffusion as a transport phenomenon contributes to oil production,
e.g. miscible displacement [83, 84] and solvent-based heavy oil recovery [80, 82, 85].
This criterion is applied in cases with competing roles of diffusion and dispersion,
i.e. convection. A high Pe implies dominant role of convection and decreased role of
molecular diffusion, such as miscible displacement. In mathematical form, the mass
transfer Pe is sometimes expressed as the product of Reynolds and Schmidt numbers
[1]:

Pe = Re · Sc (3.42)

Pe is also used to estimate dispersion coefficients [1]. When there is a high capillary
pressure in an oil reservoir, a difference between gravity and capillary forces act as a
driving force for oil flow and reservoir depletion. A second phase, gas or water, fills
the pore space during oil production and diminishes this driving force. At a balance
between capillary and gravity forces, there is no extra force available for further oil
production, although there is oil remaining in reservoir which cannot be produced
due to capillary resistance. This region is known as capillary hold-up zone. For an oil
reservoir, the ratio between capillary height, hc , and block height, L, is an important
factor that affects ultimate oil recovery. Normally, low permeability reservoirs have
higher Pc, thus remaining higher oil saturation in the capillary hold-up zone. On the
other side, reservoirs with higher permeability have lower Pc and produce more oil,
thus showing lower remaining oil saturation.

3.3 Gas Flow

The gas reservoirs are classified as conventional and unconventional. The conven-
tional gas reservoirs contain either dry gas, wet gas, or gas condensate, lean or rich
depending on the composition of intermediate components in gas. Unconventional
gas reservoirs include gas shales, coal bed methane, natural gas hydrates, and tight
3 Basics of Oil and Gas Flow in Reservoirs 101

gas reservoirs. If the reservoir gas contains CO2 and/or H2 S, the gas will be sour,
otherwise it is regarded as sweet gas when there is little or no acid gas component
in its composition. Proper evaluation of a gas reservoir performance requires such
information as gas in place, average reservoir pressure, and recoverable gas. These
are crucial information and can be extracted through reservoir simulation, and/or
production data analysis techniques. However, due to their unusual phase behavior,
reservoir studies of gas condensate fluids are associated with challenges. A detailed
methodology and analysis of challenges in gas condensate reservoirs are described
in [86–88].

3.3.1 The Diffusivity Equation for Gas Reservoirs

Flow of gas through porous media is described by a modified diffusivity equation,


in which the pressure is replaced by pseudo-pressure, pp defined by Eq. (3.43) for
radial system:

∂2 pp 1 ∂pp φμct ∂pp


+ = (3.43)
∂r 2 r ∂r 0.0002637k ∂t
Equation (3.43) is derived by assuming laminar flow, constant temperature, and
constant reservoir parameters. pp describes the relationships between PVT and trans-
port properties, compressibility factor and viscosity, with pressure in the reservoir
operating range, as is defined as Eq. (3.44):

1 2
p p ( p) = ( p − p02 ) (3.44)
μz

Details of converting general form of diffusivity equation from oil to gas systems
are given by Wattenbarger [89]. Equation (3.43) is nonlinear, as the product of
viscosity and compressibility is a function of pressure as well. However, the solution
to this nonlinear PDE is obtained similar to linear PDE developed for flow of slightly
compressible fluid like oil in porous reservoir. At certain conditions, m(P) may be
simplified to pressure squared (Eq. (3.45)) or pressure form (Eq. (3.46)):

∂ 2 p2 1 ∂ p2 φμct ∂ p2
+ = (3.45)
∂r 2 r ∂r 0.0002637k ∂t

∂2 p 1 ∂ p φμct ∂p
+ = (3.46)
∂r 2 r ∂r 0.0002637k ∂t
Equation (3.45) applies at low pressure, i.e. P < 2000 psia, where the product
(μZ) is constant. Also, Eq. (3.46) is valid at high pressure, typically P > 3000 psia,
where P/(μZ) is essentially constant. The exact pressure condition for Eqs. (3.45) and
102 R. Azin et al.

(3.46) to apply depends on gas specific gravity. The general solution of Eq. (3.46)
is similar to oil reservoir, where P is replaced by m(P). The original form of gas
diffusivity equation (Eq. (3.43)) applies to dry gas flow through porous media. In
the case of multiphase gas flow, such as gas condensate two-phase flow at pressures
between upper and lower dew-point pressure, relative permeability to each phase
and capillary pressure affect the flow pattern. Details of multiphase flow equations
are given by Ertekin et al. [90].

3.3.2 High Velocity Effects in Gas Flow

When a gas well starts production from a gas reservoir, gas flowing velocity in
reservoir increases as the stream approaches production well in a radial flow and
reaches a maximum at the wellbore. The high velocity at the wellbore implies a
transition from laminar to turbulent flow. At these conditions, the Darcy flow equation
is no longer valid. The high-velocity phenomena are characteristic of gas flow through
porous media and fractures, as well as gas flow during injection/withdrawal of gas
in underground gas storage (UGS). The high-velocity region is characteristic of gas
flow through porous media and gas flow through fractures and formation, sometimes
modelled by adding a quadratic or cubic velocity term which is frequently added to
Darcy’s law, and is commonly called the Forchheimer term [91–94]. This additional
term accounts for higher pressure drop compared to Darcy flow and is associated with
a coefficient called turbulence factor or slip factor. The cubic velocity term is used
to correct inertia term to Darcy equation for low Reynolds numbers [95]. For higher
Reynolds numbers, laminar flow switches into a transition regime and then into the
Forchheimer flow regime. This regime is described by the Forchheimer equation,
in which Darcy equation is corrected by adding a quadratic velocity term [95]. The
difference between cubic and quadratic velocity correction is that the coefficient of
cubic velocity is a combination of fluid properties only, while the quadratic velocity
correction has a coefficient composed of fluid density multiplied by rock parameter.
This parameter is known as termed the inertial resistance, β-factor, high-velocity
flow coefficient, or non-Darcy flow coefficient. The Forchheimer equation is defined
as:
dp μν
− = + βρν 2 (3.47)
dX k
The cubic form of non-Darcy equation is written as follows [96]:

dp μν
− = + βρν 2 + γρ 2 ν 3 (3.48)
dX k
Figure 3.11 show the proposed formulas for calculation of β in the literature [97].
3 Basics of Oil and Gas Flow in Reservoirs 103

Fig. 3.11 Proposed formulas for calculating β [97]

Wu et al. provided analytical solutions to the gas flow equation including Klinken-
berg effect [98]. They also proposed values for the turbulence factor and tested the
validity of the conventional assumption used for linearizing the gas flow equation.
Zeng and Grigg reviewed models and correlations for non-Darcy flow and discussed
theoretical foundations of Forchheimer number and Reynolds number [31]. They
defined non-Darcy effect (E) as the ratio of pressure gradient consumed in over-
coming liquid–solid interactions to the total pressure gradient, and Forchheimer
number, Fo, as follow:

kβρν
Fo = (3.49)
μ

βρν 2
E= (3.50)
− ddpX

Combining above equations and considering the Forchheimer equation, the


relationship between Fo and E is obtained as follows:

Fo
E= (3.51)
(1 + Fo )

Zeng and Grigg suggested that the Forchheimer number is directly connected to
the error of ignoring the non-Darcy behavior of flow in porous media [31].
104 R. Azin et al.

3.3.3 Flow Behavior of Gas Condensate Reservoirs

Any hydrocarbon reservoir experiences a pressure decline upon depletion and hydro-
carbon production. Due to its composition, the gas condensate fluid tends to become
two-phase when reservoir pressure drops below upper dew-point pressure. The inter-
mediate components of gas condensate liquefy upon pressure drop and form a film
of liquid hydrocarbon, i.e. condensate, which propagates farther into reservoir. The
condensate bank reduces gas relative permeability and well productivity. In its worst
condition, accumulation of condensate bank around wellbore may block flow of
gas. This phenomenon is referred to condensate blockage. According to Fevang and
Whitson [99], Three regions are expected to form upon gas production from a gas
condensate reservoir. The first region, just connected to wellbore contains liquid and
gas phases flowing simultaneously. The middle region has immobile liquid phase
saturation below a critical value necessary to flow. The third region contains single
phase gas and no liquid hydrocarbon saturation. The flow is single phase in regions 2
and 3, while two-phase flow prevails in region 1. Therefore, flow equations for each
region should be assigned according to the flowing phases. Extension of condensate
bank in region 2 can vary from tens of feet in lean gas condensate to hundreds of feet
in rich gas condensate reservoirs [100]. When a fluid flows through porous medium,
its momentum changes continuously as a result of changes in pore size. According to
Newton’s first law, the fluid inertia opposes momentum change and causes pressure
drop against fluid flow. This resistive force is called inertia force, which increases
with flow velocity, followed by lowering gas and liquid flow. As a result of inertia
force in two-phase flow, the relative permeability to both gas and liquid decreases.
This is known as inertia effect and may be expressed as Forchheimer coefficient,
β. The two-phase Forchheimer coefficient is used for gas-condensate systems as a
function of gas saturation and relative permeability to gas [101–103].
Similar to inertia force caused by momentum changes in flow path, viscous forces
which arise as a result of friction between fluids and pore walls act against flow and
cause pressure drop. Combination of inertia and viscous forces produces a higher
pressure drop which may result in formation of condensates in porous media. In two-
phase flow, capillary forces become active as well as viscous and inertia forces. If the
inertia forces overcome capillary forces, gas flows through bank of condensate liquid
and breaks the condensate bridge in pores. Thus, the path opens for gas and liquid
flow. This is known as positive coupling [101]. The positive coupling phenomenon
is associated with improvement in relative permeability of gas and condensate and
counteracts the negative role of inertia. Therefore, positive coupling introduces a
new concept against negative inertia in two-phase flow of gas and condensate. That
is, the two-phase relative permeability to gas and liquid is not fixed; rather, it may
increase with flow velocity as observed in long cores filled with condensate [104]. It
was further shown that the positive coupling phenomenon is a function of interfacial
properties. At large interfacial tension, the capillary force is high and phases flow in
distinct pathways in porous media. In this case, the relative permeability is mainly
a function of fluid saturation [101]. However, at low interfacial tension, i.e. near
3 Basics of Oil and Gas Flow in Reservoirs 105

miscible conditions, relative permeability depends on both liquid saturation and


interfacial tension [105].
Development of an accurate correlation is useful in predicting relative perme-
ability as a function of gas velocity, fluid saturation and interfacial tension. Such a
correlation will help in avoiding costly relative permeability curves. Different models
have been developed for relative permeability as a function of interfacial tension.
Coats proposed correlations for relative permeability of gas and condensate on the
basis of changes in interfacial tension [106]. He used an interpolating correlation
between base curves and miscible curve which approaches a straight line. Available
correlations that consider the positive coupling are obtained by two methods: The
Corey functions with the coefficients are interpolated between miscible and immis-
cible ranges, and interpolation between miscible and immiscible relative permeability
curves obtained from Coats method. Many studies support the idea that relative
permeability is a function of capillary number which depends on viscosity, velocity
and interfacial tension [105, 107, 108]. Correlations of gas/oil relative permeability
as a function of interfacial tension are also proposed by Harbert [108], Jamiolahmady
et al. [109, 110], and Fulcher et al. [111].
Also, experiments are reported which aim at the effect of interfacial tension on
relative permeability curves in gas condensate systems. Asar and Handy studied
the effect of interfacial tension on gas-oil relative permeability in gas condensate
systems [112]. They showed that relative permeability curves approach a 45° line as
interfacial tension reduces to zero. Also, by increasing interfacial tension, relative
permeability of oil (liquid phase) drops faster than relative permeability of gas.
The residual oil and gas saturation increases by increase in interfacial tension on
the other hand, at low interfacial tension near the critical value, liquid can flow at
low liquid saturation. Munkerud and Torsaeter studied the role of interfacial tension
and spreading on relative permeability of gas condensate systems [113]. They used
different fluid systems including methane, ethane, propane, pentane, hexane, and
decane with low interfacial tension and high potential for liquid drop out. Their
flow experiments showed a strong relationship between relative permeability and
gas-condensate interfacial tension. Similar experiments were reported by Henderson
et al. [114, 115], Mott et al. [103, 116], and Chen et al. [117], with a conclusion that
relative permeability of wetting phase is more sensitive to interfacial tension, while
the non-wetting phase changes slightly with decrease in interfacial tension.
On the other hand, some studies showed that interfacial tension has more influence
on relative permeability of the non-wetting phase [108, 111, 115, 118, 119]. Mean-
while, some studies suggested that relative permeability is not affected by interfacial
tension [105, 107, 112, 120, 121]. Such a contradictory finding may be due to type
of fluids and prevailing experimental conditions. In other words, it appears that the
range of experimental conditions, i.e. temperature, pressure, and composition affect
the magnitude and range of interfacial tension which may cause reverse impact on
relative permeability from a certain point.
Henderson et al. studied the role of negative inertia and positive coupling at high
flow velocity on relative permeability of gas condensate reservoirs [102]. Their results
show that inertia forces prevail in cores totally saturated with gas, while positive
106 R. Azin et al.

coupling becomes active with condensate saturation and subsequent improvement


in relative permeability was observed in all ranges of interfacial tension. According
to their experiments, positive coupling was detected at velocities up to 700 m/s and
interfacial tension up to 0.78 mN/m [102]. Jamiolahmady et al. reported the negative
inertia and positive coupling effects on gas-condensate relative permeability in low-
permeability core samples [122]. They conducted difference tests in velocity range
of 22–183 m/s and three interfacial tension values of 0.036, 0.15, and 0.85 mN/m.
Results showed that for interfacial tension of 0.85 mN/m and condensate-gas-flow
ratio (CGFR) less than 0.05, inertia effects prevail and gas relative permeability
decreases as velocity increases. Positive coupling becomes more important at higher
CGFR, resulting in relative permeability improvement. For interfacial tension of
0.15 mN/m, positive coupling was observed in velocity range of 22.8–45.7 m/s, and
inertia effect was dominant for all CGFR. For the lowest interfacial tension case
(0.036 mN/m), improvement in gas relative permeability was limited to CGFR =
0.4 and velocity range of 22.8–45.7 m/s [122].
Based on current studies, the role of positive coupling has been approved as a
mechanism of improvement in gas flow against negative inertia force which causes
additional pressure drop at high flow velocity. Approaching miscible conditions by
reducing interfacial tension between gas and condensate can highlight the role of
positive coupling. There is still much elaboration remaining to clarify the impact of
physical and transport properties like condensate API, gas and condensate viscosity,
interfacial tension, composition of gas and condensate phases, as well as operating
conditions such as pressure, temperature, gas and condensate flow rate, etc. Fruitful
research on these topics may enhance our future understanding of the exact, detailed
role of positive coupling and negative inertia in gas condensate flow through porous
media.

3.3.4 Gas Flow in Multilayer Reservoirs

The multilayer hydrocarbon formations are frequently found in nature. These forma-
tions are characterized by such features as heterogeneity, anisotropy, and sometimes
are associated with abnormal pressure profile. The gas flow in these formations
is affected by horizontal and vertical porosity and permeability variations. Gener-
ally, a dominant vertical overburden pressure implies a lower vertical permeability
compared to the horizontal permeability. In this case, the upward gas flow and migra-
tion between layers is less likely to occur, especially in the presence of impermeable
layer and each layer may be considered independent of the others. The impermeable
layers act as barriers and non-communicating layers can be studied independently
(Fig. 3.12). However, natural phenomena like tectonic stresses may generate micro
and macro fractures which can lead to excellent pathways which transfer the gas
vertically between layers.
Besides, diffusion and dispersion of gas as a result of concentration difference into
upper layers acts as another means of vertical flow. Several observations are reported
3 Basics of Oil and Gas Flow in Reservoirs 107

Fig. 3.12 Schematic of a multilayer formation with barriers in between

in the hydrocarbon reservoirs that confirm vertical gas diffusion and gradual change
in the pressure and composition of reservoirs. The most common diffusion is the
diffusion which is dependent on concentration gradients. There are other kinds of
diffusion which are dependent on pressure gradient (pressure diffusion), composi-
tion gradient (ordinary diffusion), external forces (forced diffusion) and temperature
gradient (thermal diffusion). In some situations, diffusion may compete with the
gravity forces. The gravity segregates the lighter and heavier components but, the
diffusion makes the mixture uniform [123]. Patzek et al. studied the vertical diffu-
sion in ideal gas mixtures and gravity stable in gas reservoirs. Their results show
good approximation before the concentration of diffusing gases at bottom and top
boundaries of the reservoir reach considerable concentration [124].
This type of flow between layers requires a special attention in averaging reser-
voir properties. There are three common average methods for permeability including,
weighted, harmonic and geometric average permeability. Weighted average method
is used to calculate the average permeability for some parallel layers which separate
from each other with impermeable layers. Also, each layer has its own flow rate.
But, the layers have only one flow rate in the harmonic average method. In this
method, layers face different pressure drop due to different permeability. Geometric
average permeability is introduced by Warren and Price. They reported that the
permeability of a heterogeneous reservoir could be shown by a geometric average
method. The mathematical formulation of each method and more information is avail-
able in reservoir engineering handbook by Ahmed [8]. Kuchuk and Saeedi studied
the inflow performance of horizontal wells in multi-layer reservoirs with crossflow.
They showed that if the variations of horizontal and vertical permeability are large,
the system should not be treated as a single layer with average properties. They stated
that the productivity index for an equivalent single layer system are much lower than
productivity index for actual (multi-layer) system [125]. However, in some cases, the
108 R. Azin et al.

average properties can be calculated with the Eqs. 3.52 through 3.55. These equations
could be used when the surface rate is constant and pressure behavior is assumed to
be equal to a single layer with average properties [126].
n
ki h i
k = i=1 n (3.52)
i=1 h i
n
∅i h i
∅ = i=1 n (3.53)
i=1 h i
n
i=1 C t i ki h i
Ct =  n (3.54)
i=1 ki h i
n
i=1 kr o i h i
kr o =  n (3.55)
i=1 h i

Nikjoo and Hashemi reported that three types of flow regimes could be identified
in a single phase layered reservoir with crossflow: (1) the behavior of layers without
crossflow, (2) a transition regime due to the fluid transfer between the layers, and (3)
the pressure of two layers become equal. They reported that the average properties
of equivalent single layer reservoirs can be calculated with the following equations
[127]:

h
1
k= kdz (3.56)
h
0

h
1
∅Ct = ∅Ct dz (3.57)
h
0

Fetkovich et al. stated that in the cases that crossflow exists between layers the
system could be treated as an equivalent one layer reservoir with average properties.
But, in the lack of crossflow, the multi-layer system can be combined to one equivalent
layer if: (1) the diffusivity properties of the layers are the same or (2) the ratio of flow
rate to initial gas in place is the same for each layer [128]. Park and Horne suggested
two new methods for initial estimation of the parameter’s value of a multi-layer
reservoir with crossflow. Their first method can be applied in a reservoir with two
isotropic layers. But, their second method is applicable for N layer reservoirs. More
information and mathematical description of each method are available in Park and
Horne [129].
It is important to carefully consider these phenomena in the study of gas flow in
communicating multilayer reservoirs. The question that arises is whether a multi-
layer formation can be averaged into a single reservoir with an averaged properties or
not. Dake described the methods of calculating the average of dry gas saturation and
3 Basics of Oil and Gas Flow in Reservoirs 109

pseudo-relative permeabilities for the flow of wet and dry gas in non-communicating
multilayer reservoirs for gas recycling applications. Also, he reported the mathemat-
ical expression of the vertical sweep efficiency calculation. The cumulative recovery
of wet gas (G p D ) is a function of average dry gas saturation (Sgde ), fractional flow of
dry gas ( f gde ), cumulative dry gas injected (G id ) and critical water saturation (Swc ).
Equation 3.58 shows the recovery equation reported by Dake [41]:
 
Sgde + 1 − f gde G id
G pD = (3.58)
1 − Swc

where the fractional flow for dry gas is shown as follows:

1 − 2.743 × 10−3 ϑμrgw


kk
ρ sin α
f gde =
gw
(3.59)
μgd krgw
1+ μgw krgd

It is a common practice in reservoir engineering to combine two or more layers


and consider the whole formation as a uniform layer with averaged properties. Azin
et al. studied the gas flow in a multilayer gas condensate reservoir. They comprised
the IPR curves for different number of layer. The actual reservoir has 28 production
layers. They simplified model into 1 layer and 4 layer reservoir. This simplification
can significantly decrease the run time of the model. But, the accuracy of results is an
important matter which has to be checked. Figure 3.13 shows the inflow performance
curve (IPR) for the actual reservoir and a single layer reservoir by averaging the
reservoir properties. As observed in this figure, the results of single layer reservoir
are inaccurate. The gas flow rate for a specific bottom-hole pressure (BHP) in single
layer reservoir is higher than the actual reservoir with 28 layers [130].
They observed that three layers have low permeability which can act as hinder for
pressure exchange between other layers. Also, the pressure distribution in the layers
between two impermeable layers are same because of their high vertical permeability.
They assumed that the layers between each two impermeable layer as one layer with
average reservoir properties. Figure 3.14 shows the IPR curve of actual reservoir and
the simplified four layer reservoir. As observed in this figure, the IPR curves for both
models are showed good agreement. They concluded that with this simplification the
accuracy of the model increases and also the run time of the model decreases [130].
Figure 3.15 shows the IPR and TPR curves for their studied reservoir. The inter-
section of these two curves at a given well flowing pressure and average reservoir
pressure shows the optimum operating condition. As observed from this figure, the
operating conditions of single layer model are significantly different from the 4 layers
reservoir [130].
110 R. Azin et al.

7000
Single Layer (Pr=5000 psi)
Single Layer (Pr=4000 psi)
6000 Single Layer (Pr=3000 psi)
Single Layer (Pr=2000 psi)
28-Layer (Pr=5000 psi)
Bottom-hole Pressure (psi)

5000 28-Layer (Pr=4000 psi)


28-Layer (Pr=3000 psi)
4000 28-Layer (Pr=2000 psi)

3000

2000

1000

0
0 200 400 600 800
Gas Flow Rate (MMSCFD)

Fig. 3.13 IPR curves for single layer and actual model [130]

7000
4-Layer (Pr=5000 psi)
4-Layer (Pr=4000 psi)
6000 4-Layer (Pr=3000 psi)
4-Layer (Pr=2000 psi)
28-Layer (Pr=5000 psi)
Bottom-hole Pressure (psi)

5000
28-Layer (Pr=4000 psi)
28-Layer (Pr=3000 psi)
4000 28-Layer (Pr=2000 psi)

3000

2000

1000

0
0 200 400 600 800
Gas Flow Rate (MMSCFD)

Fig. 3.14 IPR curves for 4 layers and actual model [130]
3 Basics of Oil and Gas Flow in Reservoirs 111

6000
Single Layer (Pr=5000 psi)
Single Layer (Pr=4000 psi)
Single Layer (Pr=3000 psi)
5000 Single Layer (Pr=2000 psi)
4-Layer (Pr=5000 psi)
4-Layer (Pr=4000 psi)
4-Layer (Pr=3000 psi)
Bottom-hole Pressure (psi)

4000 4-Layer (Pr=2000 psi)


TPR (Pwh=3500)
TPR (Pwh=2500)
TPR (Pwh=1500)
3000
TPR (Pwh=600)

2000

1000

0
0 100 200 300 400 500 600 700 800 900
Gas Flow Rate (MMSCFD)

Fig. 3.15 IPR and TPR curves for single and 4 layer model [130]

3.3.5 Wettability Alteration Towards Gas Wetting


of Reservoirs

As discussed above, formation and growth of condensate liquids near the wellbore
cause operation problems as well as economic disadvantages of losing valuable
condensate in reservoir. The most common methods used for eliminating condensate
blockage phenomenon include injection of gases like nitrogen, methane, propane,
carbon dioxide and gas recycling [131–136], hydraulic fracturing [137], drilling
horizontal wells [138] and injection of methanol [139] as a solvent injection. These
methods are regularly applied to remove condensate accumulation in near well-
bore region. On the other hand, these methods have become less favorable due to
high cost and lack of stability in long term. Therefore, finding an economic method
with high permanency to eliminate of this common problem is of special impor-
tance among reservoir engineers. Changing the wettability of near wellbore region
from strong liquid-wet to intermediate-wet or even strong gas-wet is a key strategy in
preventing condensate dropout and improving production from gas condensate reser-
voirs. Therefore, a recent technique which aims at controlling liquid drop out and
keeping condensate production from gas condensate reservoirs is wettability alter-
ation of near-wellbore formation towards gas wetting. Liquid repellency refers to the
ability of surface to repel liquid (such as water and oil) in the presence of other phase
112 R. Azin et al.

[140]. Knowing the fact that trapping condensate in pore spaces near the production
well causes low mobility due to the intense liquid-wetting of rock [141], several
researchers tried to investigate the effects of wettability alteration using simulation,
modelling of the process or experimental approaches.
Li and Firoozabadi [142] were the first who changed the wettability of gas conden-
sate reservoirs to gas-wet by implementing chemicals. They used fluorinated surfac-
tants to alter wettability of near wellbore region. Their results showed that perma-
nent gas-wetting state could be established in Berea sandstone and Kansas chalk via
chemical treatment. Zhang et al. [143] examined the mobility of gas and liquid phases
before and after wettability change. They showed that fluorinated polymers as wetta-
bility improvers can present a good level of water and oil repulsion from the surface of
rock, convert its wettability to intermediate gas-wet durably, increase gas and liquid
phase relative permeabilities and thereby reduce the amount of condensate blockage
and increase the productivity of wells in the reservoir condition. Also, Mousavi et al.
[141] used fluorinated nano-silica to alter wettability of limestone in the region near
wellbore of gas condensate reservoirs from strong liquid-wet to intermediate gas-wet.
Mostly, the main purpose of studies involving wettability alteration of gas condensate
reservoirs, has been finding the best chemicals to alter wettability of rock surface
durably from strong liquid-wet to gas-wet in lab scale [144, 145] and few studies
have examined the effect of wettability alteration on production factors in field scale.
One of these studies can be cited Zoghbi et al. [146]. They investigated the optimal
wettability condition to reach the maximum amount of gas condensate wells produc-
tivity through simulation process of production. Also, Sheydaeemehr et al. [147]
evaluated three different relative permeability curves that represented three states of
wettability in their model. They showed that changing wettability of reservoir rock
from liquid-wet to gas-wet and intermediate gas-wet leads to reduce condensate
saturation in near wellbore and a significant increase in condensate recovery factor.
An example of wettability alteration in field application is the research done by
Butler et al. [148], where a research team from Trueblood resources Inc., the Univer-
sity of Texas at Austin and 3M Company used chemicals for wettability alteration
purpose on a gas condensate well located in Oklahoma. Reported results showed that
near wellbore chemical stimulation could increase gas production rate by a factor of
three. In addition, their preliminary field studies indicate that wettability alteration
using chemicals can be proposed as a new effective approach while it is economically
feasible as just near wellbore is needed to be treated. The most widely used nanopar-
ticles such as TiO2 [149], Al2 O3 [150] and SiO2 [151] alter the rock wettability
from oil-wet to water-wet in water/oil system, from condensate-wet to water-wet in
condensate/water system [152], and from hydrophilic (water-wet) to hydrophobic in
water/air system. However, none of these nanoparticles can alter the contact angle
in the oil/air or liquid/air systems. The main reason can be attributed to the surface
forces, free energy and the interfacial tension between the phases and surfaces [153,
154]. Many researchers have attempted to achieve wettability alteration towards gas
wetting. Basu et al. prepared the superhydrophobic and oleophobic surface coating
by Polydimethyl siloxe. The contact angle of water and lubricant oil reached to 158–
160° and 79° on glass side, respectively. Zheng et al. studied wettability alteration of
3 Basics of Oil and Gas Flow in Reservoirs 113

glass side using silica nanoparticle modified by polytetrafluoroethylene and the result
was increased in contact angle of water to 156°. Also, the heat resistance was studied
on water contact angle [154]. Nimittrakoolchai et al. experimentally investigated the
effect of silica nanoparticle coated by trichloro (1H,1H,2H,2H-perfluorooctyl) silane
group on wettability alteration surface. Their experiments indicated water, EG and
seed oil contact angle were 173.2°, 146.7° and 147.6° [155]. Gao et al. modified
silica nanoparticle by hexadecyltrimethoxysilane group and coated on cotton and
polyester fabrics surface. Their results showed that surfaces were superhydrophobic,
so that the contact angle of cotton and polyester fabrics reached from zero to 155°
and 143°, respectively [156]. Lakshmi et al. used silica nanoparticle coated by fluo-
roalkylsilane group to study the wettability alteration. The contact angle of water,
EG, oil and n-decane have reached to 157.4°, 153.7°, 126.3° and 63.4° [157]. Li
et al. synthesized fluorinated-polyacrylate/silica core shell by emulsion polymeriza-
tion method and its effect was investigated on wettability alteration of surface [158].
They indicated that the core–shell resulted in superhydrophobic property of glass
side.
Hsieh et al. synthesized fluorinated titanium dioxide nanoparticle by chemical
vapor deposition (CVD) method [159]. They coated fluorinated titanium dioxide
nanoparticle on surface and evaluated the effect of nanoparticle on wettability alter-
ation towards hydrophilic to superhydrophobic. Also, superhydrophobic stability of
surface was investigated. Results showed that the surface was water repellency in
the presence of nanoparticle, so that the contact angle of water was about 166.1°
and the superhydrophobic stability was about 60 min. So that the surface was
superhydrophobic and oleophobic. Pazokifard et al. synthesized TiO2 nanoparticle
coated by 1H,1H,2H,2H-Perfluorooctyltriethoxysilane group [160]. They investi-
gated the effect of pH (2, 6 and 11) on synthesized nanoparticle, and concluded
that fluorosilane was adsorbed on surface of TiO2 at neutral (pH = 6) and basic
(pH = 11) conditions. Also, Yildirim et al. synthesized SiO2 nanoparticle coated by
1H,1H,2H,2H perfluorooctyltriethoxysilane group and studied its impact on wetta-
bility alteration effect of glass side [161]. The water contact angle was 161.6°.
Brassard et al. reported SiO2 nanoparticle modified by ethanolilc fluoroalkylsilane
group and coated on surface. The contact angle of water was found to be 122-
165° that showed super hydrophobicity of surface [162]. Mousavi et al. used fluo-
rinated nanosilica to change wettability from strongly condensate-wet to gas-wet
in water/condensate/gas systems, in vicinity of condensate gas wells in limestone
rock. Their results indicated change in contact angle for water and gas condensate
from zero to 124–147 and 50–70, respectively. In similar studies, Ramezani et al.
[163], Schaeffer et al. [164] Wang et al. [140], Wang et al. [165] and Valipour
Motlagh et al. [166] modified SiO2 nanoparticle by isooctyltrimethoxysilane,
tridecafluoro-1,1,2,2-tetrahydrooctyltrichlorosilane, tridecafluorooctyltriethoxysi-
lane, polytetrafluoroethylene and 1H,1H,2H,2H-perfluorodecyltriethoxysilane,
respectively [144]. Their results showed alteration of contact angle in water, EG
and fuel oil and liquid repellency of surface.
Jin et al. prepared SiO2 nanoparticle and included fluoro surfactant (FG 40) and
a polymer (FP-2) and investigated changes in contact angle of brine and n-decane.
114 R. Azin et al.

Their results indicated that contact angle of brine increased from 23° to 138°, 126°
and 150° for surfaces coated by FG40, FP-2 and modified nanosilica FG40 and the
contact angle of n-decane altered from zero to 91°, 56°, 127° coated by FG40, FP-
2 and modified nanosilica FG40, respectively. Recently, Saboori et al. synthesized
Fluorine-doped silica and Fluorine-doped silica-coating by fluorosilane nanofluid
and tested them on surface of carbonate and sandstone rock samples to study their
potential for converting surface properties towards ultrahydrophobic and ultraoleo-
phobic behavior [167]. They studied effect of nanofluid on wettability alteration and
stability of ultrahydrophobic and ultraoleophobic properties by measuring contact
angles of water, oil, condensate, n-decane and ethylene glycol in air. Their results
showed that adding 0.05 wt% of nanoparticle changes contact angle from liquid-wet
to strongly gas-wet in all systems. The original contact angle of water, oil, condensate,
n-decane and Ethylene glycol were 37.95°, 0°, 0°, 0°, 0° for carbonate rock, which
altered to 145.59°, 142.73°, 138.24°, 139.06° and 146.52° after treatment. For sand-
stone, the original contact angles were 40.40°, 0°, 0°, 0°, 0° for water, oil, condensate,
n-decane and ethylene glycol, which altered to 160.01°, 151.40°, 131.85°, 140.27°
and 151.7° after treatment [167]. They further tested novel nanoparticles on core
samples and compared results of spontaneous imbibition before and after treatment.
Imbibition of water, crude oil, and condensate into dry core changed from 0.77,
0.76 and 0.82 PV to lower than 0.10, lower than 0.10 and lower than 0.15 PV for
untreated and treated carbonate rocks [168]. In addition, the potential of synthesized
nanofluid on enhancement of gas production was studied by Sakhaei et al. [169].
They conducted contact angle, spontaneous imbibition and core flooding experi-
ments to investigate the effect of synthesized nanofluid adsorption on wettability
of rock surface and liquid mobility. Results of contact angle experiments revealed
that wettability of rock could alter from strongly oil-wetting to the intermediate
gas-wetting even at elevated temperature. Imbibition rates of oil and brine were
diminished noticeably after treatment, indicating a promising modification of rela-
tive permeability towards gas wetting and decreased condensate drop-out around
wellbore. They reported 60 and 30% enhancement in pressure drop of condensate
and brine floods after wettability alteration with modified nanofluid which confirms
successful field potential of this chemical [169].
The time to initiate wettability alteration, radius of treatment, and ultimate state of
wettability alteration, i.e. moderate or strong gas wetting are issues to be studied for a
candidate reservoir. Sakhaei et al. conducted a simulation study to figure out the role
of wettability alteration in a gas reservoir [170]. They studied effects of treatment
radius and time to optimize wettability alteration and reach the maximum condensate
production. Results indicate that near-wellbore wettability alteration leads to lower
critical condensate saturation which has a brilliant impact on improving production
parameters and reservoir recovery factors. Also, they found that highest recovery
factor in optimal condition is achieved when the wettability state of reservoir rock
is altered from strong liquid-wet to intermediate-wet, at the small radius around the
production well in early times. The strong gas-wetting, however, showed a reverse
effect and reduced the ultimate recovery of both gas and condensate. Based on this
study, wettability alteration was shown to influence inflow performance relationship
3 Basics of Oil and Gas Flow in Reservoirs 115

(IPR) curves. Curve moves downward by decreasing the average reservoir pressure,
meaning that for specific bottom-hole pressure, the well produces a lower rate at
lower average reservoir pressure. Also, by altering wettability of near wellbore zone,
the IPR curve moves upward in a given average reservoir pressure. That is to say, by
improving the wettability, the gas phase relative permeability is increased. So, gas
production rate is higher than the base case where no treatment is made. The main
reason for this favorable phenomenon is the reduction of adverse effect of condensate
blockage and pressure drop.

3.3.6 Gas Flow in Unconventional Reservoirs

Unconventional reservoirs and their distribution in the world was presented in


Chap. 1. As stated in Chap. 1, the unconventional gas reservoirs have larger reserves
in comparison with conventional ones. But, production from these kinds of reservoirs
needs specific technologies. Gas flow in unconventional reservoirs is multiscale flow
and complex process, unlike the gas flow in conventional reservoirs. The gas flow in
these reservoirs is subject to coupled processes, more non-linear, non-linear adsorp-
tion/desorption, non-Darcy flow at both high and low flow rate, rock deformation
within micro-fractures and nano-pores and strong fluid/rock interaction [171].
Shale gas reservoir is one of the unconventional gas reservoirs which consists of
large fraction of nano-pores. An apparent permeability is caused by these nano-pores.
This apparent permeability depends on fluid type, pore pressure and pore structure.
Guo et al. studied the nitrogen flow through nano membranes. Also, they constructed
a new mathematical model for characterizing the gas flow in nano pores. This model
is developed based on the advection–diffusion model. Also, they derived a new
apparent permeability based on the Knudsen diffusion and advection. Their results
show that the model can predict the experimental data with high accuracy [172].
Sun et al. reported that in the presence of organic pores smaller than 2 nm, surface
diffusion dominates the transport capacity. In contrast, the larger pore radius gives
the stronger transport capacity for inorganic pores. Humidity, stress dependence and
gas desorption affect the effective radius of nano scale pores which have significant
effects on transport capacity [173].
Slip flow, continuum flow, surface diffusion for adsorbed gas and transition flow of
bulk gas are the main mechanisms that coexist in nano pores of shale gas reservoirs.
Wu et al. developed a model by coupling an adsorbed gas surface diffusion model and
a bulk gas transport model. Their results showed that their model could describe bulk
gas transport more accurately in comparison with other models. Also, they reported
that surface diffusion has an important role in nano pores with diameter less than
2 nm (as the results reported by Sun et al. [173]). They expressed that in shale gas
reservoirs the stress dependence has high impact on fluid flow and it is related to the
effective stress, shale matrix mechanical properties and gas transport mechanisms
[174].
116 R. Azin et al.

Another kind of unconventional gas reservoir is coal-bed methane (CBM) which


contain large reserves of methane. Horizontal wells are widely used in development
of these reservoirs. The production from these reservoirs may face a period of single
phase water production followed by two phase water and gas production transient
flow. Many researchers were work on modelling the water and gas flow in CBM
reservoirs [175, 176]. Clarkson and Qanbari proposed a methodology for prediction
of under-saturated and low permeability CBM wells based on the concept of dynamic
drainage area (DDA). Their model results show a reasonable agreement with field
data [176].
Variation of coal cleat width due to coal matrix shrinkage and effective stress
which causes the variation of Klinkenberg coefficient is another challenge in recovery
process of CBM reservoirs. In most of researches, this important matter is ignored
by researchers. Wang et al. proposed two improved models, one of these models is
under constant effective stress and another is under reservoir conditions. The results
of first model show that the proportion of permeability variation is greater than the
original model due to Klinkenberg effect. Also, the Klinkenberg coefficient change
substantially however, the effect of effective stress is vanished. The second model
attached the coal porosity and apparent permeability together. The results of second
model show good agreement with field data, in particular when the gas pressure is
low [177].

Appendix 3.1: The Viscous Flow Equation in Cylindrical


Coordinates

The viscous equation is obtained in cylindrical coordinates according to Eq. (3.60)


where B = μ1 ddzP .

d 2v 1 dv
+ =B (3.60)
dr 2 r dr
This equation is an inhomogeneous ordinary differential equation because of its
non-zero right-hand side. To solve this equation, first, the homogeneous part (B = 0)
is solved and the general solution (vg ) is obtained, then the inhomogeneous part (B
= 0) is calculated, and the particular solution (vp ) is achieved, and finally, the total
answer is resulted as in Eq. (3.61):

v(r ) = vg + v p (3.61)

To solve the homogeneous part, the right-hand side of Eq. (3.60) is considered
equal to zero and the general solution (vg ) is calculated. Therefore, Eq. (3.62) is
obtained as follows:
3 Basics of Oil and Gas Flow in Reservoirs 117

d 2 vg 1 dvg
+ =0 (3.62)
dr 2 r dr
Using the variation of variables in Eq. (3.63), Eq. (3.62) rewritten as Eq. (3.64):

dvg du d 2 vg
u= . = (3.63)
dr dr dr 2
du 1
+ u=0 (3.64)
dr r
By arranging and integrating from Eq. (3.64), Eq. (3.65) is obtained as follows:

C1
u= (3.65)
r

where C 1 is a constant. By combining Eqs. (3.63) and (3.65) and reintegration,


Eq. (3.66) is achieved:

vg = C1 ln(r ) + C2 (3.66)

where C 2 is constant. Based on the point that the left-hand side of Eq. (3.60) is
a second-order differential equation, the particular solution of equation (vp ) is a
relationship in the order of two. Therefore, the particular solution of the equation
and its derivatives are obtained as Eq. (3.67):

dv p d 2v p
v p = ar 2 + br + c · = 2ar + b · = 2a (3.67)
dr dr 2

where a, b, and c are constant. Using Eq. (3.67), Eq. (3.60) can be rewritten as
Eq. (3.68):

1 b
(2a) + (2ar + b) − B = 2a + 2a + − B = 0 (3.68)
r r
Assuming b = 0 in Eq. (3.68), the relationship of constants B and a results as
Eq. (3.69):

B
a= (3.69)
4
Therefore, assuming c = 0 in Eq. (3.67), the particular solution of the differential
equation is obtained as Eq. (3.70):

B 2
vp = r (3.70)
4
118 R. Azin et al.

The overall solution of the differential equation is expressed as Eq. (3.71) using
Eqs. (3.61), (3.66) and (3.70):

B 2
v= r + C1 ln(r ) + C2 (3.71)
4
Boundary conditions are employed to calculate the constants C 1 and C 2 . Applying
the boundary conditions of Eqs. (3.72) and (3.73) in Eq. (3.71), these constants are
calculated according to Eq. (3.74):

r = ±R, v = 0 (3.72)

dv
r = 0, =0 (3.73)
dr
B 2
C1 = 0.C2 = − R (3.74)
4
Therefore, by substitution of Eq. (3.74) and B value in Eq. (3.71), the viscous
flow velocity profile equation in a cylinder can be obtained as Eq. (3.2).

1 dp  2 
v= R − r2
4μ dz

To calculate the average value of a function in cylindrical coordinates, Eq. (3.75)


is used.
R
vr dr
v = 0 R (3.75)
0 r dr

Therefore, by substituting Eq. (3.2) in Eq. (3.75), the value of the average velocity
can be calculated according to Eq. (3.76).
 R     4 
R4
1 dp 0 r R − r dr
2 2
1 dp R2 −
v= R = R2
4
(3.76)
4μ dz 4μ dz
0 r dr 2

By simplifying Eq. (3.76), the viscous flow average velocity equation in a cylinder
is obtained as Eq. (3.3).

R 2 dp
v = 0.125
μ dz
3 Basics of Oil and Gas Flow in Reservoirs 119

Appendix 3.2: The Brinkman Equation in Cylindrical


Coordinates

Brinkman equation in cylindrical coordinates is known as Eq. (3.77):

d 2v dv
r2 2
+r − Ar 2 v = r 2 B (3.77)
dr dr

In Eq. (3.77), parameter B is equal to the constant value of μ1 ddzP , and parameter A
is equal to K −1 . This equation is an ordinary inhomogeneous differential equation.
To solve this equation, both the general (vg ) and the particular (vp ) solutions which
respectively indicate the homogeneous (r 2 B = 0) and inhomogeneous part (r 2 B =
0) are calculated. According to Eq. (3.61), the sum of the particular and general
solutions is equal to the total solution of the equation. To calculate vg , first Eq. (3.77)
is homogenized as Eq. (3.78):

d 2v dv
r2 +r − Ar 2 v = 0 (3.78)
dr 2 dr
The answer of Eq. (3.78) with its similarity to the modified Basel differential
equation (reference) is obtained as Eq. (3.79):
 √   √ 
v g = C 1 I0 r A + C 2 K 0 r A (3.79)

where I 1 and K 1 are the Modified Bessel functions of the first and second kind
in the zero-order. According to the point that the left-hand side of Eq. (3.77) is a
second-order differential equation, the particular solution of equation (vp ) is in the
same order. In this case, the particular solution and its derivatives can be assumed as
Eq. (3.80):

dv p d 2v p
v p = ar 2 + br + c · = 2ar + b · = 2a (3.80)
dr dr 2

where a, b, and c are constant. Therefore, Eq. (3.77) can be rewritten as Eq. (3.81)
using Eq. (3.80).
 
r 2 (2a) + r (2ar + b) − Ar 2 ar 2 + br + c = Br 2 (3.81)

According to Eq. (3.79), it can be seen that for the validity of this equation, the
values a and b must be equal to zero (a = b = 0) and the value of c must be equal to
–B/A (c = −B/A). Therefore, by placing these constants in Eq. (3.80), the particular
solution can be calculated as Eq. (3.82).
120 R. Azin et al.

B
vg = − (3.82)
A
Finally, providing particular and general solutions and also considering Eq. (3.61),
the velocity function is calculated as Eq. (3.83).
 √   √  B
v(r ) = C1 I0 r A + C2 K 0 r A − (3.83)
A
Boundary conditions are employed to calculate the constants C 1 and C 2 in
Eq. (3.81). By applying the boundary conditions of Eq. (3.73) in Eq. (3.83), Eq. (3.84)
can be written as follows:
√  √  √  √ 
C1 AI1 0 ∗ A − C2 AK 1 0 ∗ A = 0 (3.84)

According to the Bessel function properties, the values I 1 (0) and K 1 (0) are equal
to zero and infinity, respectively. Therefore, for the validity of Eq. (3.82), the value
of C 2 must be equal to zero (C 2 = 0). Also, by applying the boundary conditions
of Eq. (3.72) and the value of C 2 in Eq. (3.83), Eq. (3.85) is achieved. By solving
Eq. (3.85), the C 1 constant results in Eq. (3.86).
 √  B
C 1 I0 R A − = 0 (3.85)
A
B
C1 =  √  (3.86)
AI0 R A

Finally, by placing the constants C 1 and C 2 as well as the values A and B in


Eq. (3.83), the velocity function is calculated as Eq. (3.87).
⎛  ⎞
I0 √r
K dP⎝
1 −  ⎠
K
v(r ) = (3.87)
μ dz I0 √RK

To calculate the average value of a function in a cylindrical system Eq. (3.75)


is used. Therefore, by placing Eq. (3.140) in Eq. (3.75), Eq. (3.88) is obtained as
follows:
⎛ R   ⎞ ⎛  √   R ⎞
R r I √r r K I √r
K dP⎝ 0 r dr 0 0 K
⎠ K dP⎜ 1 K 0 ⎟
v= R − R   = ⎝ 1 −  2 R   ⎠
μ dz R μ dz
0 r dr 0 r dr I0
√ r R
K 2
I0 √ K
0
(3.88)
3 Basics of Oil and Gas Flow in Reservoirs 121

where I 1 is the modified Bessel function of the first kind in the order of one.
By solving and simplification of Eq. (3.88), the average velocity function for the
Brinkman flow in a cylindrical system is calculated as Eq. (3.89).
⎛ √  ⎞
2 K I √R
dP K ⎝ 1
  ⎠
K
v= 1− (3.89)
dz μ R I0 √ K R

Appendix 3.3: Linear Diffusivity Equation Solution

This solution is provided by John Lee for radial flow solution in an infinite-acting
homogenous reservoir [178]. The basic partial differential equation is given in
dimensionless format

∂ 2 pD 1 ∂ pD ∂ pD
+ = (3.90)
∂r D2 r D ∂r D ∂t D

where
r
rD = (3.91)
rw
kh
p D = p DC ( pi − pr ) (3.92)
q Bμ
k
t D = t DC t (3.93)
QμCt rw2

where t DC and p DC are given by


The “initial” condition is given as

p D (r D · t D = 0) = 0 (uniform pressure distribution) (3.94)

The constant rate inner boundary condition is


 !
∂ pD
rD = −1 (constant flow rate at the well) (3.95)
∂r D r D =1

The “infinite-acting” outer boundary condition is given by

p D (r D → ∞ · t D ) = 0 (3.96)
122 R. Azin et al.

Rewriting Eq. (3.90):


 !
1 ∂ ∂ pD ∂ pD
rD = (3.97)
r D ∂r D ∂r D ∂t D

The Boltzmann transform variable, ε D , is defined as

t D = ar Db t Dc (3.98)

where in this problem


a = 1/4, b = 2, c = −1.
Which yields

r D2
εD = (3.99)
4t D

∂ 2 pD
Expanding ∂r D2

 !
∂ ∂ pD 1 ∂ pD ∂ pD
+ = (3.100)
∂r D ∂r D r D ∂r D ∂t D

Applying the chain rule

∂ pD ∂t D ∂ p D
= (3.101)
∂x ∂ x ∂t D

Which combined with Eq. (3.100) gives


 !
∂ε D ∂ ∂ε D ∂ p D 1 ∂ε D ∂ p D ∂ε D ∂ p D
+ = (3.102)
∂r D ∂ε D ∂r D ∂ε D r D ∂r D ∂ε D ∂t D ∂ε D

Expanding
   !  !
∂ε D ∂ ∂ε D ∂ p D ∂ε D ∂ 2 p D 1 ∂ε D ∂ε D ∂ p D
+ + − =0 (3.103)
∂r D ∂ε D ∂r D ∂ε D ∂r D ∂ε2D r D ∂r D ∂t D ∂t D

Isolating terms
 !2    !
∂ε D ∂ 2 pD ∂ε D ∂ ∂ε D 1 ∂ε D ∂ε D ∂ p D
+ + − =0 (3.104)
∂r D ∂ε2D ∂r D ∂ε D ∂r D r D ∂r D ∂t D ∂t D

Dividing through by (∂ε D /∂r D )2 gives


3 Basics of Oil and Gas Flow in Reservoirs 123
   !
∂ 2 pD 1 ∂ε D ∂ ∂ε D 1 ∂ε D ∂ε D ∂ p D
+ + − =0 (3.105)
∂ε2D (∂ε D /∂r D )2 ∂r D ∂ε D ∂r D r D ∂r D ∂t D ∂t D

Reducing the ∂()/∂ε D term we have


   !
∂ 2 pD 1 ∂ ∂ε D 1 ∂ε D ∂ε D ∂ p D
+ + − =0 (3.106)
∂ε2D (∂ε D /∂r D )2 ∂r D ∂r D r D ∂r D ∂t D ∂t D

Which can be further reduced to


 2 !
∂ 2 pD 1 ∂ εD 1 ∂ε D ∂ε D ∂ p D
+ + − =0 (3.107)
∂ε2D (∂ε D /∂r D )2 ∂r D2 r D ∂r D ∂t D ∂t D

Completing the factorization of the (∂ε D /∂r D )2 gives


 !
∂ 2 pD 1 ∂ 2εD 1 1 1 ∂ε D ∂ p D
+ + − =0
∂ε2D (∂ε D /∂r D )2 ∂r D2 r D ∂ε D /∂r D (∂ε D /∂r D )2 ∂t D ∂t D
(3.108)

Using Eq. (3.99) the following derivation is taken


 2   
∂ε D ∂ rD r D2 ∂ 1 −1 r D2 −1
= = = = εD (3.109)
∂t D ∂t D 4t D 4 ∂t D t D t D 4t D tD
 2 
∂ε D ∂ rD 2 2 r D2 2
= = rD = = tD (3.110)
∂r D ∂r D 4t D 4t D r D 4t D rD
  2 !  
∂ 2εD ∂ ∂ rD ∂ 2 2 2 r2 2
= = rD = = 2 D = 2 t D (3.111)
∂r D2 ∂r D ∂r D 4t D ∂r D 4t D 4t D r D 4t D rD

Substituting Eqs. (3.109)–(3.111) into Eq. (3.108) gives


  !
∂ 2 pD 1 2 1 1 1 −1 ∂ pD
+ ε D + − ε D =0
∂ε D
2
(2ε D /r D ) r D
2 2 r D (2ε D /r D ) (∂ε D /∂r D ) 2 tD ∂ε D
(3.112)

Reducing
 !
∂ 2 pD 1 1 ∂ pD
+ + + 1 =0 (3.113)
∂ε2D 2ε D 2ε D ∂ε D

Finally
 !
∂ 2 pD 1 ∂ pD
+ 1 + =0 (3.114)
∂ε2D ε D ∂ε D
124 R. Azin et al.

where Eq. (3.114) is the “Boltzmann” transformed differential equation.


Initial and boundary condition in terms of Boltzmann transform are

p D (r D · t D = 0) = 0 (3.115)

where for t D → 0; ε D → ∞, which gives

p D (ε D → ∞) = 0 (3.116)

The outer boundary condition, Eq. (3.94),

p D (r D → ∞.t D ) = 0 (3.117)

Or as r D → ∞; ε D → ∞ which yields

p D (ε D → ∞) = 0 (3.117)

where Eqs. (3.116) and (3.118) are the same which illustrates that the Boltzmann
transform “collapses” 2 conditions into 1. Combining this observation with the inner
boundary condition, we have 2 “boundary” conditions. Coupling this observation
with the fact that Eq. (3.114) is only a function of the Boltzmann variable, ε D , we
can solve Eq. (3.100) uniquely. Note that the “collapsing” of the initial and outer
boundary conditions must occur or the Boltzmann transform is technically invalid.
Recalling the constant rate inner boundary condition, Eq. (3.95)
 !  !
∂ pD ∂ pD
rD + = −1 or r D + = −1 (line source condition)
∂r D r D =1 ∂r D r D →0
(3.119)

Or
 !    !  !
∂ε D ∂ p D 2 ∂ pD ∂ pD
rD = rD εD = 2 εD + = −1
∂r D ∂ε D r D →0 rD ∂ε D r D →0,ε D →0 ∂ε D ε D →0
(3.120)

Which can be rearranged to yield


 !
∂ pD −1
εD = (3.121)
∂ε D ε D →0 2

Making the following variable of substitution

dp D
v= (3.122)
dε D
3 Basics of Oil and Gas Flow in Reservoirs 125

Substituting Eq. (3.122) into Eq. (3.124) and noting that use of ordinary derivatives
 !
dv 1
+ 1+ v=0 (3.123)
dε D εD
 !
1 1 1
dv = − 1 + dε D = −dε D − dε D (3.124)
v εD εD

Integrating

ln(v) = −ε D − ln(ε D ) + β; β = constant of integration (3.125)

Exponentiation

v = exp[−ε D − ln(ε D ) + β] (3.126)

Or
"  " 
v = exp −ε D exp − ln(ε D ) exp[β] (3.127)

Which reduces to
α
v= exp[−ε D ] (3.128)
εD

where α = exp[β], i.e., the constant of integration. Recalling Eq. (3.120) and
combining gives

dp D α
= exp[−ε D ] (3.129)
dε D εD

Multiplying through by ε D gives

dp D
εD = αexp[−ε D ] (3.130)
dε D

Substitution of Eq. (3.128) into Eq. (3.119) gives

"  −1
α lim exp[−ε D ] = (3.131)
ε D →0 2

Or

α = −1/2 (3.132)

Substitution of Eq. (3.132) into Eq. (3.129) gives


126 R. Azin et al.

dp D −1
= exp[−ε D ] (3.133)
dε D 2ε D

Separating and integrating Eq. (3.133) gives

p D ε D
−1 1 −ε D
dp D = e dε D (3.134)
2 εD
p D =0 ε D =0

where we note that p D = 0 at ε D = ∞ is the initial outer boundary condition.


Completing the integration and reversing the limits we have

∞
1 1 −y
pD = e dy (3.135)
2 y
r2
ε D = 4tD
D

We note that the integral in Eq. (3.135) is the exponential integral, E1 (x), which
is given by

∞
1 −y
E1 (x) = e dy (3.136)
y
X

Combining Eqs. (3.135) and (3.136) gives our final result


 2 
1 r
p D (r D , t D ) = E1 D (3.137)
2 4t D

Appendix 3.4: Solution of Non-linear Diffusivity Equation

The analytical solution of non-linear diffusivity equation needs changing variables


in order to perform linear equation to be solved analytically.
 2
∂2 p 1 ∂ p ∂p ∅μc ∂ p
+ + c = (3.138)
∂r 2 r ∂r ∂r kt ∂t
kt 1
p = p − pi · T = and p = ln p ∗ (3.139)
∅μc c

The linearized form is presented as [179]:


3 Basics of Oil and Gas Flow in Reservoirs 127

∂ 2 p∗ 1 ∂ p∗ ∂ p∗
+ = (3.140)
∂r 2 r ∂r ∂T
The linear equation can be solved using Laplace transform and concludes:
⎛ ⎞
      
−1 ⎝ T 0.5772λ
p = ln 1 + λ ln 2 + 0.2319 +  ⎠
c rw 1 + λ lnT + 0.2319
r 2
w

(3.141)
qμc
λ= (3.142)
4π kh

where, h is formation thickness.

Appendix 3.5: The Solution of the Warren-Root Equation


in the Different Boundary Condition

Warren-Root Equations

Reservoir pressure distribution is achieved according to the mathematical model.


Different reservoir boundary conditions can be applied to this end. The Warren-
Root model is used to calculate pressure distribution in fractured reservoirs. Equa-
tions (3.28) and (3.29) are used for fluid velocity through matrix and fractures media,
respectively:

 m = − Km grad(Pm )
U
μ

 f = − K2 grad(Pf )
U
μ

Subscripts m and f denote to matrix and fracture media, respectively. K is absolute


permeability in (m2 ), U is velocity in (ms−1 ), P is pressure in (Pa), and μ is dynamic
viscosity in (Pa s). To calculate matrix and fracture pressure distribution, the mass
conservation equation is formulated as follows:

∂(∅m ρ)  
+ div ρUm + U∗ = 0 (3.143)
∂t
∂(∅f ρ)  
+ div ρUf + U∗ = 0 (3.144)
∂t
128 R. Azin et al.

ρ SKm
U∗ = (Pm − Pf ) (3.145)
μ

The rate of mass flow per unit volume is defined by U * in (kg s−1 m−3 ), which
indicates fluid transfer between matrix and fracture in quasi-steady-state condition.
U is fluid velocity in (ms−1 ), and ϕ is porosity, which represents fluid storage capacity
in each media. S is a characteristic coefficient of fractured rock proportional to the
specific surface of a block, and ρ is fluid density in (kg m−3 ). For slightly compressible
fluids, density is calculated by Eq. (3.146).

ρ = ρ0 (1 + CP) (3.146)

where C is fluid compressibility in (Pa−1 ), and indicates dependency of fluid volume


on the pressure. Equations (3.147) and (3.148) are derived by combining Eqs. (3.143)
to (3.146).

∂Pm SKm
∅m Cm + (Pm − Pf ) = 0 (3.147)
∂t μ
  
∂Pf Kf 1 ∂ ∂Pf SKm
∅f Cf − r + (Pm − Pf ) = 0 (3.148)
∂t μ r ∂r ∂r μ

Dimensionless variables in Eqs. (3.149)–(3.153) are used to reduce the number


of variables.
2πKf h(Pi − P(r, t))
PD = (3.149)

r
rD = (3.150)
rw
Kf t
tD = (3.151)
(Cm ∅m + Cf ∅f )μr2w

αKm r2w
λ= (3.152)
Kf
∅f Cf
ω= (3.153)
∅m Cm + ∅f Cf

Dimensionless pressure is defined by PD , r D is dimensionless radius, and t D


is dimensionless time. Equations (3.152) and (3.153) represent fracture reservoir
parameters, which indicate interporosity (λ) and fracture storage capacity (ω).
At the initial time, there is a pressure equilibrium state in the reservoir. Overall
reservoir pressure is equal to reservoir initial pressure in this condition. Therefore,
at the initial times, Eq. (3.154) is used as follows:
3 Basics of Oil and Gas Flow in Reservoirs 129

P = Pi t = 0 (3.154)

Dimensionless pressure at initial time is resulted by a combination of Eqs. (3.149)


and (3.154).

PDm = 0 tD = 0 (3.155)

Equations (3.147) and (3.148) are rewritten in dimensionless forms by combining


Eqs. (3.149)–(3.153) as follows:

∂PDm
(1 − ω) − λ(PDf − PDm ) = 0 (3.156)
∂tD
 
∂PDf 1 ∂ ∂PDf ∂PDm
−ω + rD + (1 − ω) =0 (3.157)
∂tD rD ∂rD ∂rD ∂tD

Equation (3.157) is rewritten as Eq. (3.158):

∂P2Df 1 ∂PDf ∂PDm ∂PDf


+ = (1 − ω) +ω (3.158)
∂ 2 rD rD ∂rD ∂tD ∂tD

There are several methods for solving partial differential equations. Laplace trans-
form is one of these methods. Laplace transform is applied to solve differential
Eqs. (3.156) and (3.158). Consequently Eq. (3.159) is resulted by Eq. (3.156) Laplace
transform:

(1 − ω)zPDm (z, rD ) − PDm (t = 0) = λ(PDf (z, rD ) − PDm (z, rD )) (3.159)

The Laplace transform variable is known as z. By substituting Eq. (3.159) in


the initial condition (Eq. (3.155)), matrix dimensionless pressure is resulted as the
following equation:

λPDf (z · rD )
PDm (z) = (3.160)
(1 − ω)z + λ

To obtain the fracture pressure equation, the Laplace transform of Eq. (3.158) is
rewritten as Eq. (3.161).

∂ 2 PDf (z · rD ) 1 ∂PDf (z.rD )


+ = (1 − ω)zPDm (z.rD ) + ωzPDf (z.rD ) (3.161)
∂r2D rD ∂rD

Equation (3.160) is replaced in Eq. (3.161) and Eq. (3.162) is resulted as follows:

d2 PDf (z.rD ) 1 dPDf (z.rD )


+ = [zf(z)] PDf (z, rD ) (3.162)
dr2D rD drD
130 R. Azin et al.

In Eq. (3.162), f(z) is a function of ω and λ and is described as Eq. (3.31):

ω(1 − ω)z + λ
f(z) =
(1 − ω)z + λ

The solution of the partial differential Eq. (3.162) is necessary to calculate fracture
dimensionless pressure. Therefore, r D 2 is multiplied by Eq. (3.162) and Eq. (3.163)
is obtained as a standard form of Bessel equation.

d2 PDf (z.rD ) dPDf (z.rD )  2 


r2D 2
+ rD − rD zf(z) PD2 (z.rD ) = 0 (3.163)
drD drD

Therefore, the solution of Eq. (3.163) is written as Eq. (3.30) by I 0 and K 0 param-
eters, which are the first and second type of modified Bessel functions in zero-order,
respectively.

PDf (z.rD ) = C1 I0 (MrD ) + C2 K0 (MrD )

Bessel equation constants are introduced by C 1 and C 2 , and boundary √ condi-


tions are applied to obtain these values. To simplify calculations, M = zf(z) is
considered. Constants of Eq. (3.30) are calculated by different boundary conditions
substitution, and pressure distribution equations are obtained as a result.

Constant Production Rate in Closed Outer Boundary

Well production rate is kept constant in constant production rate conditions. A choke
valve is usually used to stabilize the production rate in operating conditions. This
case is more common in hydrocarbon reservoirs production. The wellbore boundary
condition is written as Eq. (3.164) in this case:
 
∂P
= constant r = rw (3.164)
∂r rw

Some hydrocarbon reservoirs are confined with faults or impermeable layers. No


fluid flows through the outer boundary in these closed reservoirs. Also, for a reservoir
with several production wells, the closed boundary can be assumed among the wells.
The closed boundary conditions are expressed in Eq. (3.165):
 
∂P
= 0 r = re (3.165)
∂r re

Darcy equation is applied to make a relationship between flow and pressure for
an inner boundary condition (the well condition) as Eq. (3.166). In this case, C 1 and
3 Basics of Oil and Gas Flow in Reservoirs 131

C 2 are calculated by replacing the boundary conditions (Eqs. (3.164) and (3.165))
in Eq. (3.30).

Kf ∂Pf
q = −2πrw h r = rw (3.166)
μ ∂r

Equation (3.165) is written for closed boundary condition and dimensionless form
of Eqs. (3.165) and (3.166) is resulted as Eqs. (3.168) and (3.167), respectively.

∂PDf
= −1 rD = 1 (3.167)
∂rD
∂PDf
= 0 rD = rDe (3.168)
∂r D

By differentiating Eq. (3.30) and integrating with Eq. (3.167), results can be
represented as the following equations:

∂PDf
C1 M I1 (M rD ) − C2 M K1 (M rD ) = (3.169)
∂rD
−1
C1 MI1 (M ) − C2 M K1 (M ) = (3.170)
z

The first and second types of modified Bessel functions in the first order are known
as I 1 and K 1 . Equation (3.171) is resulted by differentiating Eq. (3.30) and integrating
with Eq. (3.168).

C1 M I1 (M rDe ) − C2 M K1 (M rDe ) = 0 (3.171)

A system of two linear Eqs. (3.170) and (3.171) is formed to calculate C 1 and C 2 :

−K1 (M rDe )
C1 = (3.172)
z[(K1 (M rDe )) × M I1 (M ) − I1 (M rDe ) × M K1 (M )]
−I1 (MrDe )
C2 = (3.173)
z[K1 ( M rDe ) × M I1 (M ) − I1 (M rDe ) × M K1 (M )]

Equation (3.32) is resulted by substitution C 1 and C 2 in Eq. (3.30) as follows:

K1 (M rDe )I0 (M rD ) + I1 (M rDe )K0 (M rD )


PDf (z.rD ) =
zM[K1 (M)I1 (M rDe ) − K1 (M rDe )I1 (M)]

Therefore, the dimensionless pressure equation in a closed fractured reservoir and


constant production rate is given by Eq. (3.32).
132 R. Azin et al.

Constant Production Rate in Constant Pressure Outer


Boundary

For infinite reservoirs with high-pressure supplier, there is no pressure drop at the
reservoir outer boundary, and the outer boundary pressure equals to reservoir initial
pressure, as shown by Eq. (3.174). The dimensionless form of this equation is
presented as Eq. (3.175).

Pe = Pi r = re (3.174)

PDm = PDf = 0 rD = rDe (3.175)

Equation (3.175) is replaced in Eq. (3.30) and Eq. (3.176) is resulted as follows:

C1 I0 (M rDe ) + C2 K0 (M rDe ) = 0 (3.176)

A system of two linear Eqs. (3.170) and (3.176) is formed to calculate C 1 and C 2
in this case:
K0 (M rDe )
C1 = − (3.177)
z M[K0 (M rDe )I1 (M ) + I0 (M rDe )K1 (M )]
I0 (M rDe )
C2 = (3.178)
z M[K0 (M rDe )I1 (M ) + I0 (M rDe )K1 (M)]

Equation (3.33) is resulted by replacing C 1 and C 2 in Eq. (3.30).

K0 (M rD )I0 (M rDe ) − K0 (M rDe )I0 (M rD )


PDf (z.rD ) =
zM[K0 (M rDe )I1 (M) + K1 (M)I0 (M rDe )]

Equation (3.33) is applicable for dimensionless flow rate calculation at constant


pressure boundary in case of constant pressure production condition.

Constant Pressure Production in Closed Outer Boundary

The wellbore pressure is kept constant, and flow rate changes i.e., it decreases with
time in constant pressure production condition. In operating conditions, well pressure
is maintained constant by using some gauges that alter the production rate. This
production condition is mostly used in reservoirs with a high risk of gas or water
coning. The wellbore boundary condition, in this case, is written as Eq. (3.179):

P = constant r = rw (3.179)
3 Basics of Oil and Gas Flow in Reservoirs 133

Production rate equation is obtained using the wellbore pressure equation


suggested by Van Everdingen and Hurst [180]:

1
qD (z.rD ) = (3.180)
z2 PDwf (z, rD )

The dimensionless production equation (Eq. (3.34)) is obtained by combining


Eqs. (3.32) and (3.180).

M[K1 (M)I1 (M rDe ) − K1 (M rDe )I1 (M)]


qD (z.rD ) =
z[K1 (M rDe )I0 (M rD ) + I1 (M rDe )K0 (M rD )]

The dimensionless flow rate at a closed boundary fractured reservoir with constant
pressure production is calculated by Eq. (3.34).

Constant Pressure Production in Constant Pressure Outer


Boundary

In this case, the inner and outer boundary conditions are described by Eqs. (3.179)
and (3.174). The Van Everdingen and Hurst relations are applied as the same as the
previous section. Equation (3.35) is concluded by combining Eqs. (3.33) and (3.180):

M[K0 (M rDe )I1 (M) + K1 (M)I0 (M rDe )]


qD (z.rD ) =
z[K0 (M rD )I0 (M rDe ) − K0 (M rDe )I0 (M rD )]

Equation (3.35) is applicable for dimensionless flow rate calculations at constant


pressure reservoir boundary in the case of constant pressure production.

Warren-Root Analytical Solution

The original solution provided by Warren and Root considers infinite outer boundary
conditions and also the constant production rate for an inner boundary condition. The
inner boundary condition is presented as Eq. (3.164) and the outer boundary condition
is defined as Eq. (3.181):

P = Pi r = ∞ (3.181)

This assumption is applied to constant production rate conditions [65]. This


simplification is applied to Eq. (3.32). The infinite values of the first and second-order
Bessel function tend to infinite and zero, respectively. Thus, Eq. (3.36) is a simplified
134 R. Azin et al.

form of Eq. (3.32) which is resulted by using the Warren-Root assumptions:

K0 (M rD )
PD (z.rD ) =
zM K1 (M)

The Laplace inverse calculation of Eq. (3.36) is not possible by conventional


analytical methods. An approximate method for calculating K 0 and K 1 is possible
by considering primary terms of Bessel functions:
r 
D
K0 (M) = −γ − Ln M2 (3.182)
2
1
K1 (M) = (3.183)
M2
In Eq. (3.182), 7 is Euler number and equals 0.5772. This approximation is appli-
cable only for M < 0.01 values. Also, r D = 1 is assumed in the above equations.
Equation (3.184) is obtained by substituting Eqs. (3.182) and (3.183) in Eq. (3.36)
and dimensionless pressure is calculated by numerical inverse Laplace method, as
follows:
   
1 −λtD −λtD
PD (tD .1) = [0.80908 + Ln (tD ) + Ei − Ei (3.184)
2 ω(1 − ω) (1 − ω)

Exponential integral, which is called Ei function, is defined as Eq. (3.185):

∞
e−u
Ei(−X) = − du (3.185)
u
X

The Ei function values limit to zero at a very long time. Therefore, Eq. (3.184) is
converted to Eq. (3.37):

1 1
PD (tD .1) = [0.80908 + Ln(tD )] = Ln(2.25tD ) = 1.15 log(2.25tD )
2 2
Therefore, Warren and Root solved the inverse Laplace of dimensionless pressure
equation in a fractured reservoir with simple assumptions in specific conditions (M <
0.01). The solution provided by Warren and Root is a simplified form of the general
solution shown in Eq. (3.32) for M < 0.01 at wellbore (r D = 1) [63].

References

1. Bird RB, Stewart WE, Lightfoot EN. Transport phenomena, 2nd ed. (2002) n.d.
2. Brown GO. Henry Darcy and the making of a law. Water Resour Res. 2002;38:11.
3 Basics of Oil and Gas Flow in Reservoirs 135

3. Saidi AM. Realism of flow mechanism in reservoirs. 2013.


4. Løvoll G, Méheust Y, Måløy KJ, Aker E, Schmittbuhl J. Competition of gravity, capillary
and viscous forces during drainage in a two-dimensional porous medium, a pore scale study.
Energy. 2005;30:861–72.
5. Dake LP. Fundamentals of reservoir engineering, volume 8 of. Dev Pet Sci. 1978.
6. Neuman SP. Theoretical derivation of Darcy’s law. Acta Mech. 1977;25:153–70.
7. Saidi AM. Reservoir engineering of fractured reservoirs (fundamental and Practical Aspects).
Total; 1987.
8. Ahmed T. Reservoir engineering handbook. Gulf professional publishing; 2018.
9. Gavin L. Pre-Darcy flow: a missing piece of the improved oil recovery puzzle? In: SPE/DOE
symposium on improved oil recovery. Society of Petroleum Engineers; 2004.
10. Amyx J, Bass D, Whiting RL. Petroleum reservoir engineering physical properties. 1960.
11. Ahmed T, McKinney P. Advanced reservoir engineering. Elsevier; 2011.
12. Klinkenberg LJ. The permeability of porous media to liquids and gases. In: Drilling and
production practice. American Petroleum Institute; 1941.
13. Azin R, Mohamadi-Baghmolaei M, Sakhaei Z. Parametric analysis of diffusivity equation in
oil reservoirs. J Pet Explor Prod Technol. 2017;7:169–79.
14. Basak P. Non-Darcy flow and its implications to seepage problems. J Irrig Drain Div.
1977;103:459–73.
15. Soni JP, Islam N, Basak P. An experimental evaluation of non-Darcian flow in porous media.
J Hydrol. 1978;38:231–41.
16. Xue-wu W, Zheng-ming Y, Yu-ping S, Xue-wei L. Experimental and theoretical investigation
of nonlinear flow in low permeability reservoir. Procedia Environ Sci. 2011;11:1392–9.
17. Siddiqui F, Soliman MY, House W, Ibragimov A. Pre-Darcy flow revisited under experimental
investigation. J Anal Sci Technol. 2016;7:1–9.
18. Prada A, Civan F. Modification of Darcy’s law for the threshold pressure gradient. J Pet Sci
Eng. 1999;22:237–40.
19. Hekecioglu I, Jiang Y. Flow through porous media of packed spheres saturated within water.
J Fluids Eng. 1994;116:164–70.
20. Civan F. Porous media transport phenomena. Wiley; 2011.
21. Evans RD, Civan F. Characterization of non-Darcy multiphase flow in petroleum bearing
formation. Final report. Oklahoma Univ., Norman, OK (United States). School of Petroleum
and …; 1994.
22. Dye AL, McClure JE, Miller CT, Gray WG. Description of non-Darcy flows in porous medium
systems. Phys Rev E. 2013;87:33012.
23. Mitchell J, Younger J. Abnormalities in hydraulic flow through fine-grained soils. Permeability
and capillarity of soils. ASTM International; 1967.
24. Yu RZ, Bian YN, Lei Q, Yang ZM, Wang KJ. Low-velocity non-Darcy flow numer-
ical simulation of the Da45 reservoir, Honggang Oilfield, Jilin, China. Pet Sci Technol.
2013;31:1617–24.
25. Farmani Z, Azin R, Fatehi R, Escrochi M. Analysis of pre-Darcy flow for different liquids
and gases. J Pet Sci Eng. 2018;168:17–31.
26. Farmani Z, Farokhian D, Izadpanahi A, seifi F, Zahedizadeh P, Safari Z, et al. Pre-Darcy
flow and Klinkenberg effect in dense, consolidated carbonate formations. Geotech Geol Eng.
2019;37. https://doi.org/10.1007/s10706-019-00841-0.
27. Brinkman HC. A calculation of the viscous force exerted by a flowing fluid on a dense swarm
of particles. Flow, Turbul Combust. 1949;1:27.
28. De Lemos MJS. Turbulence in porous media: modeling and applications. Elsevier; 2012.
29. Dybbs A, Edwards RV. A new look at porous media fluid mechanics—Darcy to turbulent.
Fundamentals of transport phenomena in porous media. Springer; 1984. p. 199–256
30. Hubbert MK. Darcy’ law: its physical theory and application to entrapment of oil and gas.
History of geophysics: volume 3: the history of hydrology, vol. 3. 1987. p. 1–26.
31. Zeng Z, Grigg R. A criterion for non-Darcy flow in porous media. Transp Porous Media.
2006;63:57–69.
136 R. Azin et al.

32. Zimmerman RW, Al-Yaarubi A, Pain CC, Grattoni CA. Non-linear regimes of fluid flow in
rock fractures. Int J Rock Mech Min Sci. 2004;41:163–9.
33. Hassanizadeh SM, Gray WG. High velocity flow in porous media. Transp Porous Media.
1987;2:521–31.
34. Zahedizadeh P, Safari Z, Behnamnia M, Sadeqi-Murche-khorti R, Azadian H, Mirghayed BR,
et al. Boundary determination between darcy, brinkman, and viscous flows using effective
parameters. Appl Math Comput 2021;Under Revi.
35. Terry RE, Rogers JB, Craft BC. Applied petroleum reservoir engineering. Pearson Education;
2015.
36. Mosavat N, Torabi F, Zarivnyy O. Developing new Corey-based water/oil relative perme-
ability correlations for heavy oil systems. In: SPE heavy oil conference. Society of Petroleum
Engineers; 2013.
37. Torabi F, Mosavat N, Zarivnyy O. Predicting heavy oil/water relative permeability using
modified Corey-based correlations. Fuel. 2016;163:196–204.
38. Sufi AH, Ramey Jr HJ, Brigham WE. Temperature effects on relative permeabilities of oil-
water systems. In: SPE annual technical conference and exhibition. Society of Petroleum
Engineers; 1982.
39. Honarpour MM. Relative permeability of petroleum reservoirs. CRC press; 2018.
40. Xu J, Guo C, Jiang R, Wei M. Study on relative permeability characteristics affected
by displacement pressure gradient: experimental study and numerical simulation. Fuel.
2016;163:314–23.
41. Dake LP. The practice of reservoir engineering. vol. 36. Elsevier; 2013.
42. Mohamadi-Baghmolaei M, Azin R, Sakhaei Z, Mohamadi-Baghmolaei R, Osfouri S. Novel
method for estimation of gas/oil relative permeabilities. J Mol Liq. 2016;224:1109–16.
43. Purcell WR. Capillary pressures-their measurement using mercury and the calculation of
permeability therefrom. J Pet Technol. 1949;1:39–48.
44. Fatt I, Dykstra H. Relative permeability studies. J Pet Technol 1951;3:249–56.
45. Burdine N. Relative permeability calculations from pore size distribution data. J Pet Technol.
1953;5:71–8.
46. Corey AT. The interrelation between gas and oil relative permeabilities. Prod Mon.
1954;19:38–41.
47. Wyllie MRJ, Gardner GHF. The generalized Kozeny-Carman equation. Part 2. A novel
approach to problems of fluid flow. World Oil 1958;146:210–28.
48. Wahl WL, Mullins LD, Elfrink EB. Estimation of ultimate recovery from solution gas-drive
reservoirs. Trans AIME. 1958;213:132–8.
49. Torcaso MA, Wyllie MRJ. A comparison of calculated krg/kro ratios with a correlation of
field data. J Pet Technol. 1958;10:57–8.
50. Brooks RH, Corey AT. Properties of porous media affecting fluid flow. J Irrig Drain Div.
1966;92:61–90.
51. Honarpour M, Koederitz LF, Harvey AH. Empirical equations for estimating two-phase
relative permeability in consolidated rock. J Pet Technol. 1982;34:2–905.
52. Ibrahim MN, Koederitz LF. Two-phase relative permeability prediction using a linear
regression model. In: SPE eastern regional meeting. Society of Petroleum Engineers; 2000.
53. Mulyadi H, Amin R, Kennaird AF. Practical approach to determine residual gas saturation and
gas-water relative permeability. In: SPE annual technical conference and exhibition. Society
of Petroleum Engineers; 2001.
54. Lomeland F, Ebeltoft E, Thomas WH. A new versatile relative permeability correlation. Paper
SCA 2005-32 presented at international symposium of the Society of Core Analysts, Toronto,
Canada; 2005.
55. Sigmund PM, McCaffery FG. An improved unsteady-state procedure for determining the
relative-permeability characteristics of heterogeneous porous media (includes associated
papers 8028 and 8777). Soc Pet Eng J. 1979;19:15–28.
56. Chierici GL, Ciucci GM, Sclocchi G. Two-phase vertical flow in oil wells-prediction of
pressure drop. J Pet Technol. 1974;26:927–38.
3 Basics of Oil and Gas Flow in Reservoirs 137

57. Chierici GL. Novel relations for drainage and imbibition relative permeabilities. Soc Pet Eng
J. 1984;24:275–6.
58. Bastian P. Numerical computation of multiphase flow in porous media, Habilitation. Tech Fak
Der Christ Kiel. 1999.
59. Li K, Horne RN. Comparison of methods to calculate relative permeability from capillary
pressure in consolidated water-wet porous media. Water Resour Res 2006;42.
60. Ahmadi MA, Zendehboudi S, Dusseault MB, Chatzis I. Evolving simple-to-use method to
determine water–oil relative permeability in petroleum reservoirs. Petroleum. 2016;2:67–78.
61. Ahmadi MA. Connectionist approach estimates gas–oil relative permeability in petroleum
reservoirs: application to reservoir simulation. Fuel. 2015;140:429–39.
62. Farmani Z, Azin R, Mohamadi-Baghmolaei M, Fatehi R, Escrochi M. Experimental and
theoretical study of gas/oil relative permeability. Comput Geosci. 2019;23:567–81.
63. van Golf-Racht TD. Fundamentals of fractured reservoir engineering, vol. 12. Elsevier; 1982.
64. Barenblatt GI, Zheltov IP, Kochina IN. Basic concepts in the theory of seepage of
homogeneous liquids in fissured rocks [strata]. J Appl Math Mech. 1960;24:1286–303.
65. Warren JE, Root PJ. The behavior of naturally fractured reservoirs. Soc Pet Eng J. 1963;3:245–
55. https://doi.org/10.2118/426-PA.
66. Zahedizadeh P, Safari Z, Ghaderi A, Azin R. Transition zone diagnosis in finite fractured
reservoirs. Comput Geosci. 2020;24:1483–96.
67. Dong Z, Li W, Lei G, Wang H, Wang C. Embedded discrete fracture modeling as a method
to upscale permeability for fractured reservoirs. Energies. 2019;12:812.
68. Choi ES, Cheema T, Islam MR. A new dual-porosity/dual-permeability model with non-
Darcian flow through fractures. J Pet Sci Eng. 1997;17:331–44.
69. Hoteit H, Firoozabadi A. Multicomponent fluid flow by discontinuous Galerkin and mixed
methods in unfractured and fractured media. Water Resour Res 2005;41.
70. Mousavi Nezhad M, Javadi AA, Abbasi F. Stochastic finite element modelling of water flow in
variably saturated heterogeneous soils. Int J Numer Anal Methods Geomech. 2011;35:1389–
408.
71. Karimi-Fard M, Durlofsky LJ, Aziz K. An efficient discrete fracture model applicable for
general purpose reservoir simulators. In: SPE reservoir simulation symposium proceedings.
Society of Petroleum Engineers; 2003.
72. Fumagalli A, Pasquale L, Zonca S, Micheletti S. An upscaling procedure for fractured
reservoirs with embedded grids. Water Resour Res. 2016;52:6506–25.
73. Ding M, Kantzas A. Capillary number correlations for gas-liquid systems. J Can Pet Technol
2007;46.
74. Kiani M, Osfouri S, Azin R, Dehghani SAM. Impact of fluid characterization on compositional
gradient in a volatile oil reservoir. J Pet Explor Prod Technol. 2016;6:835–44.
75. Kiani M, Osfouri S, Azin R, Dehghani SAM. A comprehensive model for the prediction of
fluid compositional gradient in two-dimensional porous media. J Pet Explor Prod Technol.
2019;9:2221–34.
76. Azin R, Raad SMJ, Osfouri S, Fatehi R. Onset of instability in CO 2 sequestration into
saline aquifer: scaling relationship and the effect of perturbed boundary. Heat Mass Transf.
2013;49:1603–12.
77. Raad SMJ, Fatehi R, Azin R, Osfouri S, Bahadori A. Linear perturbation analysis of
density change caused by dissolution of carbon dioxide in saline aqueous phase. J Mol Liq.
2015;209:539–48.
78. Louriou C, Ouerfelli H, Prat M, Najjari M, Nasrallah S Ben. Gas injection in a liquid satu-
rated porous medium. Influence of gas pressurization and liquid films. Transp Porous Media
2012;91:153–71.
79. Saidi AM. Twenty years of gas injection history into well-fractured Haft Kel field (Iran). In:
International petroleum conference and exhibition of Mexico. Society of Petroleum Engineers;
1996. https://doi.org/10.2118/35309-MS.
80. Alkindi AS, Muggeridge A, Al-Wahaibi YM. The influence of diffusion and dispersion on
heavy oil recovery by VAPEX. In: International thermal operations and heavy oil symposium.
Society of Petroleum Engineers; 2008.
138 R. Azin et al.

81. Azin R, Kharrat R, Ghotbi C, Rostami B, Vossoughi S. Simulation study of the VAPEX
process in fractured heavy oil system at reservoir conditions. J Pet Sci Eng. 2008;60:51–66.
82. Etminan SR, Haghighat P, Maini BB, Chen ZJ. Molecular diffusion and dispersion coef-
ficient in a propane-bitumen system: case of vapour extraction (VAPEX) process. In: SPE
EUROPEC/EAGE annual conference and exhibition. Society of Petroleum Engineers; 2011.
83. Adewale A, Olalekan O, Kelani B, Abass I. Effects of Peclet number on miscible displacement
process through the reservoir. In: Nigeria annual international conference and exhibition.
Society of Petroleum Engineers; 2004.
84. Patel RD, Greenkorn RA. Prediction of recovery in miscible displacement in porous media
using the dispersion equation. In: Fall meeting of the Society of Petroleum Engineers of
AIME. Society of Petroleum Engineers; 1969.
85. Boustani A, Maini BB. The role of diffusion and convective dispersion in vapour extraction
process. J Can Pet Technol 2001;40.
86. Sureshjani MH, Azin R, Gerami S. A roadmap for production data analysis in an Iranian
offshore gas-condensate field n.d.
87. Heidari Sureshjani MH, Azin R, Lak A, Osfouri S, Chahshoori R, Sadeghi F, et al. Production
data analysis in a gas-condensate field: methodology, challenges and uncertainties. Iran J
Chem Chem Eng. 2016;35:113–27.
88. Sarvestani MT, Sureshjani MH, Gerami S, Azin R. Some remarks on production data analysis
in layered gas reservoirs. n.d.
89. Wattenbarger RA. Well performance equations (1987 PEH Chapter 35). Pet Eng Handb. 1987.
90. Ertekin T, Abou-Kassem JH, King GR. Basic applied reservoir simulation. 2001.
91. Firoozabadi A, Katz DL. An analysis of high-velocity gas flow through porous media. J Pet
Technol. 1979;31:211–6.
92. Tek MR, Coats KH, Katz DL. The effect of turbulence on flow of natural gas through porous
reservoirs. J Pet Technol. 1962;14:799–806.
93. Ezeudembah AS, Dranchuk PM. Flow mechanism of Forchheimer’s cubic equation in high-
velocity radial gas flow through porous media. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1982.
94. Holditch SA, Morse RA. The effects of non-Darcy flow on the behavior of hydraulically
fractured gas wells (includes associated paper 6417). J Pet Technol. 1976;28:1–169.
95. Skjetne E, Auriault J-L. High-velocity laminar and turbulent flow in porous media. Transp
Porous Media. 1999;36:131–47.
96. Eburi Losoha S. Mathematical modeling, analysis and simulation of the productivity index
for non-linear flow in porous media, with applications in reservoir engineering. 2007.
97. Saboorian-Jooybari H, Pourafshary P. Significance of non-Darcy flow effect in fractured tight
reservoirs. J Nat Gas Sci Eng. 2015;24:132–43.
98. Wu Y-S, Pruess K. Gas flow in porous media with Klinkenberg effects. Transp Porous Media.
1998;32:117–37.
99. Whitson CH, Fevang O. Modeling gas condensate well deliverability, SPE 30714. 1995.
100. Fan L, Harris BW, Jamaluddin AJ. Understanding gas-condensate reservoirs. Oilf Rev.
2005:14–27. https://doi.org/10.4043/25710-MS.
101. Jamiolahmady M, Danesh A, Tehrani DH, Duncan DB. A mechanistic model of gas-
condensate flow in pores. Transp Porous Media. 2000;41:17–46.
102. Henderson GD, Danesh A, Tehrani DH, Al-Kharusi B. The relative significance of positive
coupling and inertial effects on gas condensate relative permeabilities at high velocity. In:
SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2000.
103. Mott R, Cable A, Spearing M. Measurements and simulation of inertial and high capillary
number flow phenomena in gas-condensate relative permeability. In: SPE annual technical
conference and exhibition. Society of Petroleum Engineers; 2000.
104. Danesh A, et al. Gas condensate recovery studies. 1994.
105. Longeron DG. Influence of very low interfacial tensions on relative permeability. Soc Pet Eng
J. 1980;20:391–401.
106. Coats KH. SPE 8284. An equation of state compositional model. 1980.
3 Basics of Oil and Gas Flow in Reservoirs 139

107. Amaefule JO, Handy LL. The effect of interfacial tensions on relative oil/water permeabilities
of consolidated porous media. Soc Pet Eng J. 1982;22:371–81.
108. Harbert LW. Low interfacial tension relative permeability. In: SPE annual technical conference
and exhibition. Society of Petroleum Engineers; 1983.
109. Jamiolahmady M, Danesh A, Tehrani DH, Duncan DB. Positive effect of flow velocity on
gas–condensate relative permeability: network modelling and comparison with experimental
results. Transp Porous Media. 2003;52:159–83.
110. Jamiolahmady M, Sohrabi M, Ireland S, Ghahri P. A generalized correlation for predicting
gas–condensate relative permeability at near wellbore conditions. J Pet Sci Eng. 2009;66:98–
110.
111. Fulcher RA Jr, Ertekin T, Stahl CD. Effect of capillary number and its constituents on two-
phase relative permeability curves. J Pet Technol. 1985;37:249–60.
112. Asar H, Handy LL. Influence of interfacial tension on gas/oil relative permeability in a gas-
condensate system. SPE Reserv Eng. 1988;3:257–64.
113. Munkerud PK, Torsaeter O. The effects of interfacial tension and spreading on relative perme-
ability in gas condensate systems. In: IOR 1995—8th European symposium on improved oil
recovery. European Association of Geoscientists & Engineers; 1995. p. cp-107.
114. Henderson GD, Danesh A, Tehrani DH, Al-Shaidi S, Peden JM. Measurement and correlation
of gas condensate relative permeability by the steady-state method. SPE J. 1996;1:191–202.
115. Henderson GD, Danesh A, Tehrani DH, Al-Shaidi S, Peden JM. Measurement and correlation
of gas condensate relative permeability by the steady-state method. SPE Reserv Eval Eng.
1998;1:134–40.
116. Mott R, Cable A, Spearing M. A new method of measuring relative permeabilities for calcu-
lating gas-condensate well deliverability. In: SPE annual technical conference and exhibition.
Society of Petroleum Engineers; 1999.
117. Chen HL, Wilson SD, Monger-McClure TG. Determination of relative permeability and
recovery for North Sea gas condensate reservoirs. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1995.
118. Blom SMP, Hagoort J, Soetekouw DPN. Relative permeability at near-critical conditions. In:
SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1997.
119. Henderson GD, Danesh A, Tehrani DH, Peden JM. The effect of velocity and interfacial
tension on relative permeability of gas condensate fluids in the wellbore region. J Pet Sci Eng.
1997;17:265–73.
120. Delclaud J, Rochon J, Nectoux A. Investigation of gas/oil relative permeabilities: high-
permeability oil reservoir application. In: SPE annual technical conference and exhibition.
Society of Petroleum Engineers; 1987.
121. Kalaydjian F-M, Bourbiaux BJ, Lombard JM. Predicting gas-condensate reservoir perfor-
mance: how flow parameters are altered when approaching production wells. In: SPE annual
technical conference and exhibition. Society of Petroleum Engineers; 1996.
122. Jamiolahmady M, Sohrabi M, Ghahri P, Ireland S. Gas/condensate relative permeability of a
low permeability core: coupling vs. inertia. SPE Reserv Eval Eng 2010;13:214–27.
123. Thomas O. Reservoir analysis based on compositional gradients. 2007.
124. Patzek TW, Silin DB, Benson SM, Barenblatt GI. On vertical diffusion of gases in a horizontal
reservoir. Transp Porous Media. 2003;51:141–56.
125. Kuchuk FJ, Saeedi J. inflow performance of horizontal wells in multilayer reservoirs. In: SPE
annual technical conference and exhibition. Society of Petroleum Engineers; 1992.
126. Jatmiko W, Daltaban TS, Archer JS. Multi-phase flow well test analysis in multi-layer reser-
voirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers;
1996.
127. Nikjoo E, Hashemi A. Pressure transient analysis in multiphase multi layer reservoirs with
inter layer communication. In: SPE Europec/EAGE annual conference. Society of Petroleum
Engineers; 2012.
128. Fetkovich MJ, Bradley MD, Works AM, Thrasher TS. Depletion performance of layered
reservoirs without crossflow. SPE Form Eval. 1990;5:310–8.
140 R. Azin et al.

129. Park H, Horne RN. Well test analysis of a multilayered reservoir with formation crossflow.
In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1989.
130. Azin R, Sedaghati H, Fatehi R, Osfouri S, Sakhaei Z. Production assessment of low production
rate of well in a supergiant gas condensate reservoir: application of an integrated strategy. J
Pet Explor Prod Technol. 2019;9:543–60.
131. Ahmed T, Evans J, Kwan R, Vivian T. Wellbore liquid blockage in gas-condensate reservoirs.
In: SPE eastern regional meeting. Society of Petroleum Engineers; 1998.
132. Jamaluddin AKM, Ye S, Thomas J, D’Cruz D, Nighswander J. Experimental and theoretical
assessment of using propane to remediate liquid buildup in condensate reservoirs. In: SPE
annual technical conference and exhibition. Society of Petroleum Engineers; 2001.
133. Kossack CA, Opdal ST. Recovery of condensate from a heterogeneous reservoir by the injec-
tion of a slug of methane followed by nitrogen. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1988.
134. Marokane D, Logmo-Ngog AB, Sarkar R. Applicability of timely gas injection in gas conden-
sate fields to improve well productivity. In: SPE/DOE improved oil recovery symposium.
Society of Petroleum Engineers; 2002.
135. Narinesingh J, Alexander D. CO2 enhanced gas recovery and geologic sequestration in
condensate reservoir: a simulation study of the effects of injection pressure on condensate
recovery from reservoir and CO2 storage efficiency. Energy Procedia. 2014;63:3107–15.
136. Sanger PJ, Hagoort J. Recovery of gas-condensate by nitrogen injection compared with
methane injection. SPE J. 1998;3:26–33.
137. Wang X, Indriati S, Valko PP, Economides MJ. Production impairment and purpose-built
design of hydraulic fractures in gas-condensate reservoirs. In: International oil and gas
conference and exhibition in China. Society of Petroleum Engineers; 2000.
138. Dehane A, Tiab D, Osisanya SO. Comparison of the performance of vertical and hori-
zontal wells in gas-condensate reservoirs. In: SPE annual technical conference and exhibition.
Society of Petroleum Engineers; 2000.
139. Asgari A, Dianatirad M, Ranjbaran M, Sadeghi AR, Rahimpour MR. Methanol treatment
in gas condensate reservoirs: a modeling and experimental study. Chem Eng Res Des.
2014;92:876–90.
140. Wang H, Chen E, Jia X, Liang L, Wang Q. Superhydrophobic coatings fabricated with polyte-
trafluoroethylene and SiO2 nanoparticles by spraying process on carbon steel surfaces. Appl
Surf Sci. 2015;349:724–32.
141. Mousavi MA, Hassanajili S, Rahimpour MR. Synthesis of fluorinated nano-silica and its
application in wettability alteration near-wellbore region in gas condensate reservoirs. Appl
Surf Sci. 2013;273:205–14.
142. Kewen L, Abbas F. Experimental study of wettability alteration to preferential gas-wetting in
porous media and its effects. SPE Reserv Eval Eng. 2000;3:139–49.
143. Zhang S, Jiang G-C, Wang L, Qing W, Guo H-T, Tang X, et al. Wettability alteration to
intermediate gas-wetting in low-permeability gas-condensate reservoirs. J Pet Explor Prod
Technol. 2014;4:301–8.
144. Jin J, Wang Y, Wang K, Ren J, Bai B, Dai C. The effect of fluorosurfactant-modified nano-silica
on the gas-wetting alteration of sandstone in a CH4-liquid-core system. Fuel. 2016;178:163–
71.
145. Li K, Liu Y, Zheng H, Huang G, Li G. Enhanced gas-condensate production by wettability
alteration to gas wetness. J Pet Sci Eng. 2011;78:505–9.
146. Zoghbi B, Fahes MM, Nasrabadi H. Identifying the optimum Wettability conditions for the
near-wellbore region in gas-condensate reservoirs. In: Tight gas completions conference.
Society of Petroleum Engineers; 2010.
147. Sheydaeemehr M, Sedaeesola B, Vatani A. Gas-condensate production improvement using
wettability alteration: a giant gas condensate field case study. J Nat Gas Sci Eng. 2014;21:201–
8.
148. Butler M, Trueblood JB, Pope GA, Sharma MM, Baran Jr JR, Johnson D. A field demon-
stration of a new chemical stimulation treatment for fluid-blocked gas wells. In: SPE annual
technical conference and exhibition. Society of Petroleum Engineers; 2009.
3 Basics of Oil and Gas Flow in Reservoirs 141

149. Ehtesabi H, Ahadian MM, Taghikhani V, Ghazanfari MH. Enhanced heavy oil recovery in
sandstone cores using TiO2 nanofluids. Energy Fuels. 2014;28:423–30.
150. Giraldo J, Benjumea P, Lopera S, Cortés FB, Ruiz MA. Wettability alteration of sandstone
cores by alumina-based nanofluids. Energy Fuels. 2013;27:3659–65.
151. Hendraningrat L, Torsæter O. Effects of the initial rock wettability on silica-based nanofluid-
enhanced oil recovery processes at reservoir temperatures. Energy Fuels. 2014;28:6228–41.
152. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate
cores by alumina-nanofluid in different base fluids and temperature. J Sustain Energy Eng.
2018;6:84–98.
153. Basu BJ, Kumar VD, Anandan C. Surface studies on superhydrophobic and oleophobic poly-
dimethylsiloxane–silica nanocomposite coating system. Appl Surf Sci. 2012;261:807–14.
154. Zheng Y, He Y, Qing Y, Zhuo Z, Mo Q. Formation of SiO2/polytetrafluoroethylene hybrid
superhydrophobic coating. Appl Surf Sci. 2012;258:9859–63.
155. Nimittrakoolchai O-U, Supothina S. Preparation of stable ultrahydrophobic and superoleo-
phobic silica-based coating. J Nanosci Nanotechnol. 2012;12:4962–8.
156. Gao Q, Zhu Q, Guo Y, Yang CQ. Formation of highly hydrophobic surfaces on cotton and
polyester fabrics using silica sol nanoparticles and nonfluorinated alkylsilane. Ind Eng Chem
Res. 2009;48:9797–803.
157. Lakshmi RV, Bera P, Anandan C, Basu BJ. Effect of the size of silica nanoparticles on wetta-
bility and surface chemistry of sol–gel superhydrophobic and oleophobic nanocomposite
coatings. Appl Surf Sci. 2014;320:780–6.
158. Li K, Zeng X, Li H, Lai X. Fabrication and characterization of stable superhydrophobic
fluorinated-polyacrylate/silica hybrid coating. Appl Surf Sci. 2014;298:214–20.
159. Hsieh C-T, Lai M-H, Cheng Y-S. Fabrication and superhydrophobicity of fluorinated titanium
dioxide nanocoatings. J Colloid Interface Sci. 2009;340:237–42.
160. Pazokifard S, Mirabedini SM, Esfandeh M, Farrokhpay S. Fluoroalkylsilane treatment of
TiO2 nanoparticles in difference pH values: Characterization and mechanism. Adv Powder
Technol. 2012;23:428–36.
161. Yildirim A, Budunoglu H, Daglar B, Deniz H, Bayindir M. One-pot preparation of fluorinated
mesoporous silica nanoparticles for liquid marble formation and superhydrophobic surfaces.
ACS Appl Mater Interfaces. 2011;3:1804–8.
162. Brassard J-D, Sarkar DK, Perron J. Synthesis of monodisperse fluorinated silica nanoparticles
and their superhydrophobic thin films. ACS Appl Mater Interfaces. 2011;3:3583–8.
163. Ramezani M, Vaezi MR, Kazemzadeh A. Preparation of silane-functionalized silica films via
two-step dip coating sol–gel and evaluation of their superhydrophobic properties. Appl Surf
Sci. 2014;317:147–53.
164. Schaeffer DA, Polizos G, Smith DB, Lee DF, Hunter SR, Datskos PG. Optically trans-
parent and environmentally durable superhydrophobic coating based on functionalized SiO2
nanoparticles. Nanotechnology. 2015;26:55602.
165. Wang H, Fang J, Cheng T, Ding J, Qu L, Dai L, et al. One-step coating of fluoro-containing
silica nanoparticles for universal generation of surface superhydrophobicity. Chem Commun.
2008:877–9.
166. Motlagh NV, Birjandi FC, Sargolzaei J, Shahtahmassebi N. Durable, superhydrophobic, super-
oleophobic and corrosion resistant coating on the stainless steel surface using a scalable
method. Appl Surf Sci. 2013;283:636–47.
167. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Synthesis of fluorine-doped silica-
coating by fluorosilane nanofluid to ultrahydrophobic and ultraoleophobic surface. Mater Res
Express. 2017;4:105010.
168. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate
rocks from strongly liquid-wetting to strongly gas-wetting by fluorine-doped silica coated by
fluorosilane. J Dispers Sci Technol. 2018;39:767–76.
169. Sakhaei Z, Azin R, Naghizadeh A, Osfouri S, Saboori R, Vahdani H. Application of fluori-
nated nanofluid for production enhancement of a carbonate gas-condensate reservoir through
wettability alteration. Mater Res Express. 2018;5:35008.
142 R. Azin et al.

170. Sakhaei Z, Mohamadi-Baghmolaei M, Azin R, Osfouri S. Study of production enhance-


ment through wettability alteration in a super-giant gas-condensate reservoir. J Mol Liq.
2017;233:64–74.
171. Wu Y-S, Li J, Ding D, Wang C, Di Y. A generalized framework model for the simulation of
gas production in unconventional gas reservoirs. SPE J. 2014;19:845–57.
172. Guo C, Xu J, Wu K, Wei M, Liu S. Study on gas flow through nano pores of shale gas
reservoirs. Fuel. 2015;143:107–17.
173. Sun Z, Li X, Shi J, Zhang T, Sun F. Apparent permeability model for real gas transport
through shale gas reservoirs considering water distribution characteristic. Int J Heat Mass
Transf. 2017;115:1008–19.
174. Wu K, Chen Z, Li X, Guo C, Wei M. A model for multiple transport mechanisms through
nanopores of shale gas reservoirs with real gas effect–adsorption-mechanic coupling. Int J
Heat Mass Transf. 2016;93:408–26.
175. Sun Z, Shi J, Wang K, Miao Y, Zhang T, Feng D, et al. The gas-water two phase flow behavior
in low-permeability CBM reservoirs with multiple mechanisms coupling. J Nat Gas Sci Eng.
2018;52:82–93.
176. Clarkson CR, Qanbari F. A semi-analytical method for forecasting wells completed in low
permeability, undersaturated CBM reservoirs. J Nat Gas Sci Eng. 2016;30:19–27.
177. Wang G, Ren T, Wang K, Zhou A. Improved apparent permeability models of gas flow in
coal with Klinkenberg effect. Fuel. 2014;128:53–61.
178. Lee J. Well testing. Society of Petroleum Engineers; 1982.
179. Odeh AS, Babu DK. Comparison of solutions of the nonlinear and linearized diffusion
equations. SPE Reserv Eng. 1988;3:1–202.
180. Van Everdingen AF, Hurst W. The application of the Laplace transformation to flow problems
in reservoirs. J Pet Technol. 1949;1:305–24. https://doi.org/10.2118/949305-G.
Chapter 4
Gas Injection for Underground Gas
Storage (UGS)

Reza Azin and Amin Izadpanahi

Abstract In this chapter, some of the key aspects of UGS as a sustainable energy
supply infrastructure were reviewed. This type of gas injection is associated with the
reuse of depleted oil and gas formations as a natural underground storage volumes
with proved reservoir characteristics and cap rock integrity, as well as available
surface facilities and wells. The operation has its distinct features in terms of planning,
gas injection/withdrawal, and etc. Some of the common criteria for screening the UGS
were also reviewed and discussed in this chapter.

Abbreviations

Acronyms

CHR Condensate holding ratio


CPE Condensate production enhancement
GE Gas equivalent
I/W Injection/withdrawal
OGIP Original gas in place

R. Azin (B)
Faculty of Petroleum, Gas and Petrochemical Engineering, Department of Petroleum Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 143
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_4
144 R. Azin and A. Izadpanahi

Variables

B Formation volume factor, reservoir conditions/standard conditions


bpss Pressure loss due to the steady-state inflow of gas, psia/ft3 /day
c Compressibility, 1/psi
c1 Matrix pore volume compressibility, 1/psi
c2 Fracture pore volume compressibility
ce Effective compressibility, 1/psi
G Gas volume, scf
h Net pay thickness, ft
K K-Value
k Permeability, md
Mw Molecular weight, g/gmol
N Condensate volume, STB
P Average drainage area pressure, psia
P Pressure, psia
PD Dimensionless pressure
PDi Dimensionless pressure integral derivative
Pp Pseudopressure, psia2 /cp
Ppwf Dimensionless bottom-hole pseudopressure
Pr Pressure in the reservoir at the drainage radius, psia
Pw Bottomhole pressure, psia
q Flow rate, ft3 /day
Qp Production rate of a given phase, SCF/D
Rep Drainage radius, ft
re Exterior radius, ft
rw Wellbore radius, ft
S Saturation
T Temperature, F
tca Material balance pseudotime for gas, day
tDA Dimensionless time
tp Production time, day
V Volume, acre-ft
W Water Volume, bbl
x Liquid phase mole fraction
y Gas phase mole fraction
Z Average gas compressibility factor
Z Gas compressibility factor

Greek Letters

φ Porosity
4 Gas Injection for Underground Gas Storage (UGS) 145

φ1 Matrix void space to bulk volume


φ2 Fracture void space to bulk volume
γ Specific gravity
μ Viscosity

Subscripts

b Bulk
e Influx
f Formation
g Gas
i Initial
inj Injected
ng Non-gas
p Produced
sc Standard condition
so Sour gas injection
sw Sweet gas injection
t Total
w Water

4.1 Introduction

Natural gas has now become one of the world’s main energy source in only less than
two decades. According to international energy prospect scenario, the global natural
gas consumption during years 2001–2025 will experience an average growth rate of
2.9–3.2% per year which is comparable to an annual growth rate of 1.8% for oil and
1.5% for coal [1]. Countries must plan and consider further investments to guarantee
a constant rate of natural gas supply for the coming years.
Worldwide resources of natural gas and the need for clean and low GHG emission
fuel has made a shift towards consumption of natural gas. Underground gas storage
(UGS) projects are becoming an essential part of gas and market chain in countries to
resolve seasonal demand and mitigate pick shaving problems efficiently. UGS plays
a crucial role in energy management, balancing demand and supply chain as well as
regulating the gas price. This chapter will review the key features of UGS technology
as a special application of gas injection.
146 R. Azin and A. Izadpanahi

4.2 The Concept of UGS

Underground gas storage (UGS) is defined as the storage of large quantities of natural
gas in a storage formation in order to support the natural gas demand in domestic,
commercial, industrial, or space heating as the most crucial application especially in
winter. The strong trend towards increasing the number of UGS facilities is observed
in gas importing and hub countries an economical, safe and environment-friendly
technology operating tool.
Generally, UGS includes two steps of injection and withdrawal (I/W) in each
cycle. In the injection phase, natural gas is injected into subsurface formation when
market demand drops below the supply available from the pipeline. Natural gas is
then withdrawn out of storage reservoir to provide a steady supply from the pipeline
when demand exceeds supply.
Some UGS reservoirs are designed and operated to provide seasonal gas to desig-
nated and predictably constant markets. They are called base load types of reservoirs.
Other UGS projects are designed to respond only at extreme demand for gas, called
peak shaving reservoirs. There are three basic requirements attributed to the perfor-
mance of UGS reservoirs, including verification of inventory, assurance of deliver-
ability, and containment against migration [2]. The inventory presents existing gas
in the storage reservoir. It is made up of base gas (or cushion gas) and top gas (or
working gas). Deliverability refers to the ability of UGS field to deliver gas to its
dedicated market and depends on the equalized pressure prevailing underground.
Since the pressure is a function of the amount of gas in the storage container, it
follows that deliverability is a function of inventory. If the container does not hold
the gas, it becomes subject to the attrition of its inventory through the migration of
gas.
Based on the source of gas and final destination of stored gas, the UGS may be
operated in the production zone, hub, or consumption zone. The UGS in produc-
tion zone involves long-term gas storage into depleted gas reservoirs located in the
gas production zone. This type of gas injection is of special importance for rapid
development and depletion of gas reservoirs shared between two or more countries.
Unless there is a valid agreement between holding countries, shared reservoirs are
subject to depletion by each country independently. At these conditions, there is a
possibility for unbalanced gas production and gas migration as a consequence. This
unbalanced production may be a result of a delay in the development of production
facilities, pipeline, surface and processing facilities, gas boosting stations, unbal-
anced gas consumption, and etc. At these conditions, one solution for balancing
gas withdrawal from a shared reservoir will be UGS in a production zone. Another
option for UGS is storing gas in geological formations located in export hubs. For
example, countries like Ukraine act as a hub for gas export between Russia and the
rest of European countries. These UGS hubs store gas volumes in larger amounts
required by the country itself, and the UGS is designed mainly for gas export. The
third scenario for UGS is storing gas in underground formations to compensate peak
demands in the consumption and populated areas. This scenario necessitates UGS in
4 Gas Injection for Underground Gas Storage (UGS) 147

formations near the consumption areas and is known as the UGS at the consumption
zone.
The UGS is a mature technology both in surface facilities and underground. Yet,
there are uncertain features in cap-rock integrity, reservoir inventory, well design, and
etc. when UGS is planned for a new reservoir.

4.3 History of UGS in the World

The first experimental project of UGS was conducted in Welland County, Ontario,
Canada in 1915 [3]. One year later in 1916, the first operational UGS project on the
bases of depleted field was created in Zoar field near Buffalo, New York, USA. The
first UGS project in the aquifer was operated in 1946 in Kentucky, USA. In 1961,
the first UGS project in salt cavern was started in Michigan, USA, and the first UGS
operation in an abandoned mine and lined rock cavern were 1963 in Colorado, USA
and 2002 in Skallen, Sweden [4]. The idea of storing natural gas in underground
reservoirs during low-consumption for use in high-demand seasons and meet the
peak rates has found worldwide application since 1950s.

4.4 Candidate Reservoirs

The most common type of gas storage projects are operated in underground reser-
voirs, classified in three principal types, i.e. depleted gas reservoirs, aquifer reser-
voirs, un-minable coal beds and salt cavern reservoirs. Most early storage pools and a
high percentage of all pools today were developed in depleted reservoirs [5]. Besides
these types of UGS, Tureyen et al. describe other means of gas storage, i.e. storage in
pipelines, storage in high pressure steel reservoirs, storage as liquefied natural gas,
and storage in manmade caverns [5].
Malakooti and Azin describe the main features, advantages and disadvantages
of three main UGS approaches. Storage in salt caverns is quite different than in a
depleted oil and gas reservoir and aquifer. The salt cavern is initially washed by water
and a large cave is formed underground to ensure high deliverability of gas during
peak season. Whereas the gas flow occurs through porous media in the case UGS
in depleted reservoir and saline aquifer. The key properties of a UGS plan can be
summarized as follow [3]:
1. Supply of natural gas flow for gas consumers is guaranteed especially during
the cold season for peak shaving
2. Natural gas provided for gas consumers when technical problems occur in gas
processing facilities
3. Balance between natural gas production and consumption is established during
all time
148 R. Azin and A. Izadpanahi

4. Significant decrease in required investments to build new capacities for gas


transportation system of the country
5. Establish a strategic resource for emergency and critical situations like war,
earthquake, and etc.
6. Increased withdrawal from shared gas fields and reinjection into domestic
reservoirs
7. Providing necessary facilities in business markets such as buying and storing
cheap natural gas for domestic consumption, obtaining higher income by
importing stored gas at high gas price increases in market.
A real case study of a UGS project involves subsurface and surface design and
engineering studies. The surface facility for UGS is similar to other gas injection
applications and will be discussed in Chap. 7. The subsurface studies include all
aspects of geosciences, i.e. geology, geophysics, petrophysics, reservoir engineering,
and production engineering. From the reservoir engineering aspect, the following
parameters are important:
• Determine reservoir storage capacity and reservoir performance
• Optimize well design, well number and well location in the UGS reservoir
• Compare the performance of horizontal versus vertical well in the UGS reservoir
• Analyze well productivity and deliverability
• Design injection/withdrawal rate scenarios
• Determine oil/condensate recovery in UGS operated in depleted oil or gas
reservoir
• Determine the role of impurities like acid gas (CO2 /H2 S) in the injecting stream
on reservoir and produced gas compositions.
• Perform and evaluate the reservoir material balance.

4.5 Worldwide UGS Projects

Nowadays, underground gas storage projects have become an essential part of


the gas chain due to supply-demand unbalances of natural gas globally, more gas
consumption in the world, and more usage of underground reservoirs to store natural
gas.
Statistical studies show that 432 underground reservoirs have been designed for
UGS plans in 11 countries by 1974, 300 of which were located in the United
States. Based on reports by CEDIGAZ Company, the total number of 554 UGS
reservoirs were operating worldwide in 1993. These storage reservoirs were able to
deliver 243 MMMm3 working gas which amounts to 11% of world gas usage in
1993. In 2009, the number of underground storage facilities increased to 630 with
352.48 MMMm3 working gas. Statistical studies illustrate that UGS plans last aver-
agely eight to ten years for each reservoir. UGS projects represent the fact that a
production capacity of 100 MMMm3 is installing in the world. This capacity will
not be enough because 2% growth in gas consumption increases the value of gas
4 Gas Injection for Underground Gas Storage (UGS) 149

Table. 4.1 Number of active


Region Number of active projects
projects based on the region
USA 392
Canada 62
Europe 143
Common Wealth of Independent 48
States (CIS)
Asia-Oceania 23
Middle East 3
Argentina 1

consumption up to 3100 MMMm3 in 2010. CEDIGAZ predicted that a capacity


equal to 200 MMMm3 is required with the investment of $b40. At the end of 2016,
672 UGS facilities were in operation worldwide, representing a working gas capacity
of 424 bcm equal to 12% of 2016 world gas consumption. Table 4.1 shows the UGS
projects based on the region. Global UGS capacity is expected to increase from
413 bcm in 2015 to between 547 and 640 bcm in 2035, which requires an investment
of $b123–270 [6].
Recognition of appropriate underground structures to store gas has started in Iran
since 1989. Due to high consumption pattern, average daily gas consumption in four
months of winter is about 2.5 times that in the rest of year in Iran. This rate may reach
4 times during the extreme cold days of winter. Consequently, supplementary sources
of gas are needed in peak demand season which can be provided by UGS. NIGC
established an independent management called UGS company with the mission to
develop UGS projects in 2005. Currently, there is a UGS facility operating in Sarajeh
formation, Iran. Sarajeh is an anticline reservoir located 40 km far from south east
of Qom, about 160 km south of Tehran. The structure sizes are 25 × 5 km. This
field has a storage capacity of 3.3 MMMm3 gas. Subsurface studies determined that
7.3 MMm3 gas should be injected into Sarajeh field during the summer season. After
that 9.8 MMm3 gas should be produced from the reservoir during four cold months of
the year. It was established if the reservoir performance satisfied the results obtained
by researches, the injection and production rates will increase up to 22 and 33 MMm3
per day, respectively [3].

4.6 Screening Criteria for UGS

Generally, UGS reservoirs consist of good to excellent quality formations which


are often located spatially close to the ultimate demand source (i.e. major popula-
tion centers). Most of these reservoirs represent natural gas pools which have been
depleted near to or below abandonment pressure during normal production oper-
ations, then are used on a seasonal basis for gas storage. For a reservoir to be a
candidate for gas storage, the following criteria must be satisfied [7]:
150 R. Azin and A. Izadpanahi

1. Sufficient reservoir volume allows for storage of the required amount of gas
without exceeding containment pressure constraints and without requiring
uneconomic compression to high pressure levels.
2. Satisfactory containment of the gas by upper and lower sealing cap rock.
3. Sufficient inherent permeability to allow injection and production at required
delivery rates during peak demand periods.
4. Limited sensitivity to reductions in permeability, well injectivity/productivity
due to the presence of in-situ water (mobile or immobile), presence of liquid
hydrocarbons (mobile or immobile), plugging of the near injector region by
compressor lubricants or other introduced fluids, and reservoir stress fluctuations
during successive pressure cycles.
5. Absence of hydrogen sulfide gas (in-situ or bacterially generated).
6. Ability to drill and complete additional wells in the formation as required with
causing severe formation damage (due to the highly depleted pressure condition
which may often exist in these reservoirs).
Gas storage reservoirs are generally high permeability clastics or carbonates
(1000–10,000 mD in-situ permeability is common) existing at intermediate depths
and temperatures. In general, these reservoirs are depleted formations which orig-
inally contained dry (non-retrograde), sweet (no H2 S) natural gas. Typically, these
zones do not contain mobile water or active or partially active aquifers, oil legs or
residual liquid hydrocarbon saturation, although this is not always the case [7].
Another issue which must be considered in reservoir screening is yearly cyclic
gas injection-shut-in-withdrawal pattern. According to Fig. 4.1, a complete cycle
of UGS involves injection, intermediate shut-in, withdrawal, and final shut-in. The
intermediate shut-in period lets the gas to evenly distribute in reservoir and relax
the whole reservoir before withdrawal. In terms of reservoir inventory, the UGS
is theoretically expected to move on a linear pattern on P/z versus Gp plot for a
volumetric reservoir. However, there are frequently disturbances observed in UGS
pressure profile indicating deviation from linear pattern. According to Fig. 4.1, a
shut-in period between successive injection and withdrawal may relax deviations in

Fig. 4.1 A complete cycle of UGS involves injection, intermediate shut-in, withdrawal, and final
shut-in
4 Gas Injection for Underground Gas Storage (UGS) 151

reservoir pressure. Normally, a 0.5–2 month shut-in time is suggested by simulation


studies [8] and field operations. In a UGS operation in Iran, a stable foam produc-
tion was observed which severely affected the surface facility, especially separators,
instruments, and flow meters, so that the operator decided anti-foam injection at
early cycles of I/W. In next cycles, the foaming phenomenon was greatly resolved
by extending the intermediate shut-in period and letting the reservoir to relax before
withdrawal.
Figure 4.2 shows the pressure behavior of a UGS cycle and its relationship with
the reservoir inventory. The ideal general material balance equation suggests a linear
pressure trend versus gas in place shown by dash line X–Y. However, the real field
operation is expected to follow a cycle like ABCD in an aquifer-supported reservoir
and A B C D volumetric reservoir. The UGS cycle starts from point A at the start
of injection season and the reservoir is charged with injected gas up to point B,
where a shut-in time causes a slight reduction in the measured pressure. During
the shut-in time, pressure distribution becomes uniform throughout a reservoir with
a permeability pattern typical of a sandstone formation. Therefore, the reservoir
pressure at the end of injection is expected to be at point C for an aquifer-supported
reservoir and point C for a volumetric reservoir. For a tight formation with low

1600

1500
B Y
B'
Reservoir with Aquifer C
1400 C'
Volumetric Reservoir

1300
Reservoir Pressure (psia)

1200

1100

1000

900

A'
800 A
D
700 D'
X

600
12 14 16 18 20 22 24 26 28
Total Gas in Place (Bcf)

Fig. 4.2 Comparison of cycles in volumetric reservoir with a reservoir with aquifer
152 R. Azin and A. Izadpanahi

permeability, the difference between pressure at points B and C is small. Also, The
difference between points C and C implies the potential of aquifer to support the
reservoir pressure. The withdrawal stage starts from point C and ends at point D (D
for a volumetric reservoir). Again, the reservoir pressure becomes uniform during a
shut-in time and point D approaches point A at the end of withdrawal. In an ideal
reservoir with uniform rock properties, the cycle may approach the X–Y line. This
figure also suggests that part of the pressure decline at the end of withdrawal stage
is restored by the aquifer and the pressure at point D is higher than point D for a
volumetric reservoir. The degree of communication between reservoir and aquifer
determines the difference between pressures at the end of injection and withdrawal
for two reservoirs.
Table 4.2 summarizes the screening criteria for UGS obtained from field opera-
tions. The formation thickness, depth, porosity, permeability, reservoir pressure and
temperature are among the important factors that should be taken into account for
UGS planning and selecting the candidate formations. In the studied cases, most
formations were depleted oil reservoirs. Two formations were gas condensate, one
gas reservoir, and a rock salt.

4.7 UGS Key Aspects

4.7.1 Threshold Pressure

The minimum pressure required for displacing the cap-rock pore water is defined as
the threshold pressure. First, the gas enters the cap rock but it doesn’t have the ability
to flow through the cap rock. After continuous flow to the cap rock, the gas saturation
exceeds the critical value. In this stage, the gas flows through the cap rock. Threshold
pressure can measure the sealing ability of cap rock. This characteristic is directly
measured in lab experiments which has different procedures and approaches. Also,
some theoretical relations has been developed based on the rock properties such as
porosity, permeability and formation resistivity for predicting the threshold pressure
[19].

4.7.2 Well Testing

The well testing for UGS project needs some specific requirements in comparison
to the conventional well testing. Repeated evaluation must be done to predict the
formation damage due to the high risk of clogging of gravel pack or screen and
sand production. Also, well productivity and injectivity is tested during the well
testing for characterizing the effect of turbulence flow on well performance during
the storage cycle. The turbulence flow in this stage is caused by the significant
Table. 4.2 Screening criteria of some UGS projects and studies
Reservoir fluid Field Formation Formation depth Porosity Permeability Reservoir pressure Reservoir Refs
thickness (ft) (ft) (md) (psi) temperature (F)
Oil field Pecorade NM 8202 0.06 0.05 3843 210.2 [9]
Oil field IZ-2 Reservoir NM NM 0.25 NM 2260 140 [10]
Gas field Niger Delta NM 1860 0.354 NM 816 119 [11]
Oil field New York City 300 6100 0.155 NM 3214.7 142 [12]
Gas condensate NM* 2200 6300 Low Low NM NM [13]
Gas field Sarajeh 5800 0.065 20 5699 228 [14]
Oil field Triassic 229.6 NM 0.037 50 4235 NM [15]
Hauptdolomite
4 Gas Injection for Underground Gas Storage (UGS)

Weathered Zone 164 NM 0.062 90


Bockfleis 65.6 NM 0.085 400
Formation
Rock salt Jintan Salt Mine 656 3260 NM NM NM NM [16]
Oil field Sockeye Field 70 5600 0.235 20–300 2494 NM [17]
Gas condensate Tiehchenshan 164 9022 0.18 Ph * = 50–300, 3938 223 [18]
Pv * = 25–150
NM = Not mentioned
Ph = Horizontal permeability
Pv = Vertical permeability
153
154 R. Azin and A. Izadpanahi

variation of reservoir pressure. Identification of the lateral and vertical variations of


reservoir petrophysical properties are another important subject in well testing of
a UGS project. These variations can cause uneven pressure distribution. Dynamic
response of the system can be affected by minor heterogeneities due to high rates
constrained by storage operations [19].

4.7.3 Rock Mechanics

Extension and magnitude of fault stability and ground movements are evaluated with
geo-mechanical analyses. Geo-mechanical behavior of the system is characterized
under existing reservoir conditions by in-situ measurements and lab experiments on
cores. Three most important properties of rock mechanics are in-situ stresses, intact
rocks and faults. More details of this section can be found in Verga et al. [19].

4.8 Reservoir Fluid Phase Behavior in UGS

PVT study in UGS reservoirs depends on the type of reservoir under study. For
depleted oil and gas reservoirs, PVT study involves characterization of original fluid
remaining in place at the end of depletion, selection and tuning of a proper EOS
using detailed PVT experiments to match the experimental data. This approach is
reported in the works of Khamehchi and Rashidi [14], Xiao et al. [4], Azin et al.
[3, 20–22]. Details of PVT study techniques were reviewed in Chap. 2. In the case
of UGS in depleted gas condensate reservoir, condensate revaporization may occur
which results in fluctuations of CGR in successive I/W cycles [21]. For UGS in saline
aquifers, solubility of gas into aquifer and water content of injected gas upon contact
with reservoir water must be determined either through experiments or predicted by
models. Zirrahi et al. proposed exact models for prediction of water content [23] and
solubility of gas in saline aquifer [24]. Azin et al. reported that substantial changes
in compressibility factor of reservoir fluid occurs in successive I/W cycles [25]. This
is due to sharp compositional changes in reservoir fluid as a result of contact and
mixing with injection gas.

4.8.1 Injection Gas Versus Reservoir Fluid

During UGS, the injected gas contacts reservoir fluid and mixes with it. Depending
on the composition of injection and reservoir fluids, extent of contact, time of contact,
and level of mixing, the phase behavior of reservoir fluid is subject to change. The
cyclic injection of dry gas can vaporize the light and intermediate components of
4 Gas Injection for Underground Gas Storage (UGS) 155

reservoir oil. Also, it may re-vaporize the condensate bank in gas-condensate reser-
voirs. However, the radius and depth of these effects are limited in large reservoirs
and it takes time for the reservoir boundary to “sense” the effect of gas storage. The
effects of injected gas on the reservoir fluid phase behavior was discussed in Chap. 2.

4.8.2 Sour Gas Versus Sweet Gas

In addition to sweet gas injection, sour gas, a blend of natural gas, hydrogen sulfide
(H2 S) and carbon dioxide (CO2 ) streams, can be considered as another option for
the purpose of UGS in gas production hub and in formations with sufficient gas
treating facilities. Doing so can increase the gas withdrawal from shared reservoirs
and increase the strategic reserves of country without the need to gas treatment.
Underground sweet gas storage causes double costs in the industry, as sour gas
stream is treated before injection and after withdrawal to remove contamination
with fluid remaining in reservoir. Storage of sour gas is a more economic option
for UGS projects because the sweetening and treating process is performed once.
In this manner, the rate of gas withdrawal would be higher from shared reservoirs
since operators do not need to construct treatment facilities before reinjection of the
produced gas.
When UGS is implemented in depleted gas condensate reservoirs, it can enhance
condensate recovery through gas injection. Injection of dry gas increases the dew
point pressure and helps revaporization of heavy hydrocarbon fractions into gas
phase [26]. Pure CO2 and separator gases containing H2 S and CO2 are common
injection streams to minimize condensate drop-out in near-wellbore zones [27]. A
decrease in the compressibility factor is another impact of acid gases present in the
injection gas. Based on compositional simulation, CO2 is found to be an efficient dry
gas stream that can be injected into the retrograde reservoir to improve condensate
recovery [28]. Changes in the wetting behavior of reservoir rock and interfacial
tension between the storage gas and reservoir fluid are other important issues during
UGS. Melean et al. [29] investigated the influence of gas/condensate interfacial
tension and wetting behavior of condensate on the porous substrate. They found
that CO2 is more advantageous than CH4 and N2 in reducing capillary forces and
promoting the spreading of the condensate phase on water.
Sour natural gas mixtures cause problems like corrosion in wellbore and surface
facilities, and increased tendency for hydrate formation at elevated pressures [30].
These issues imply that all aspects of sour gas mixtures including their advantages
and disadvantages must be considered before commencing UGS using sour gas.
Malakooti and Azin compared the sweet gas injection with three sour gas streams.
The compositions of each stream are reported in Table 4.3. The sweet gas stream
contains no CO2 and H2 S. Sour gas stream 3 has the highest mole fraction of H2 S
[20].
156 R. Azin and A. Izadpanahi

Table. 4.3 Compositions of the injected gas streams [20]


Composition Sweet gas Sour gas stream 1 Sour gas stream 2 Sour gas stream 3
C1 –N2 0.975 0.955116 0.906784 0.851784
C2 –CO2 0.0246 0.034498 0.07742 0.12742
C3 –nC4 0.0004 0.006978 0.008613 0.008613
iC4 –nC5 0 0.000045 0.000047 0.000047
FC6 0 0.000001 0.000001 0.000001
C7 –C11 0 0 0 0
C13+ 0 0.000012 0.000013 0.000013
H2 S 0 0.00335 0.007122 0.012122

0.97
Sweet Gas Stream
Sour Gas Stream 1
0.95 Sour Gas Stream 2
Sour Gas Stream 3
0.93
Z Factor

0.91

0.89

0.87

0.85
0 1000 2000 3000 4000
Pressure (psia)

Fig. 4.3 Gas compressibility factor versus pressure [20]

Figure 4.3 shows the gas compressibility factor of different gas streams. Obvi-
ously, the presence of CO2 and H2 S components in the injection gas stream decreases
the compressibility z-factor of the injection stream [20].
Figure 4.4 indicates the difference between the compressibility factor of reservoir
fluid and injected gas. The difference between the compressibility z-factor of injected
gas and reservoir fluids is smaller in the case of sour gas streams. Also, decreasing
trend of each stream is because of decreasing the difference between the compositions
of reservoir fluid and injected gas [20].
Figure 4.5 shows the pressure rise at the end of I/W cycle. This figure indicates
that by increasing the mole fraction of CO2 and H2 S the reservoir pressure at the end
of I/W cycle will be decreased. This fact is another reason of the smaller difference
between the reservoir and injected gas compressibility factor [20].
Figure 4.6 shows the ratio of difference between heating value (HV) of produced
gas and injected gas to HV of the injected gas. This figure illustrates the effect of
H2 S on the HV of produced gas. This ratio is increased by increasing the mole
percent of H2 S. A stream with 0% of H2 S mole fraction has the lowest difference
4 Gas Injection for Underground Gas Storage (UGS) 157

0.03
Sweet Gas Stream
Sour Gas Stream 1
0.025 Sour Gas Stream 2
Sour Gas Stream 3
0.02
Zinj-Zres

0.015

0.01

0.005

0
0 2 4 6 8 10
I/W Cycle

Fig. 4.4 Difference between the compressibility factor of reservoir fluid and injected gas during
I/w cycles [20]

2400
Sweet Gas Stream
2390 Sour Gas Stream 1
Sour Gas Stream 2
2380 Sour Gas Stream 3
Pressure (psia)

2370

2360

2350

2340

2330

2320
1 3 5 7 9
I/W Cycle

Fig. 4.5 Pressure at the end of I/W cycle [20]

between the injected and produced gas in comparison to other streams. Also, heating
value of produced gas increases during each I/W cycle due to the increasing of
heavy component mole fraction in produced gas. Heavy component mole fraction in
gas phase increases due to the pressure decline of the reservoir. As the total heavy
component of the reservoir decreases, the production of heavy components is also
decreased. As a results, the heating value of production gas stream decreases [20].
Figure 4.7 demonstrates the heating value for injection and production streams
with different percentage of H2 S. The sweet gas stream has the highest heating value.
By increasing the H2 S percent, the heating value is decreased in both injection and
production streams [20].
158 R. Azin and A. Izadpanahi

25

100*(HVprod-HVinj)/HVinj (%)
Sweet Gas
20 Sour Gas 5% H2S
Sour Gas 10% H2S

15

10

0
0 2 4 6 8 10
I/W Cycle

Fig. 4.6 Ratio of difference between HV of produced and the injected gas streams to HV of injected
gas stream [20]

1150
Prod (Sweet Gas)
Prod (Sour Gas 5% H2S)
1100
Heating Value (Btu/Cu ft)

Prod (Sour Gas 10% H2S)


Inj (Sweet Gas)
Inj (Sour Gas 5% H2S)
1050 Inj (Sour Gas 10% H2S)

1000

950

900

850
1 2 3 4 5 6 7 8 9 10 11
I/W Cycle

Fig. 4.7 HV of production and injection streams versus successive I/W cycles [20]

4.8.3 Condensate Re-vaporization During UGS

Liquid condensate formed by retrograde condensation during primary depletion in


gas condensate reservoirs is trapped in porous media due to its low saturation. Injec-
tion of pipeline gas into this reservoir increases pressure and cause trapped condensate
to evaporate and mix with the injected and residual gases. The withdrawn gas will
be a mixture of injected gas, evaporated condensate and residue gas. As a result, the
withdrawn gas will contain heavy hydrocarbons and must be processed to prevent
condensation in the gathering and transmission distribution systems. It should be
4 Gas Injection for Underground Gas Storage (UGS) 159

noted that significant volume of liquid can be produced from the withdrawn gas even
though the liquid yield during storage withdrawal cycle is significantly lower than the
liquid yield during the primary production. This is mainly due to high gas withdrawal
rates during storage operation. Proper design of surface and processing facilities to
handle the produced liquid is essential for reliable operation of the storage field.
Condensate yield depends on the composition of withdrawn gas from storage. The
degree of mixing among residue gas, evaporated condensate, and the injected gas
determines the withdrawn gas composition and depends on several factors including
well configuration, the residence time of injected gas, reservoir structure, and I/W
schedule.
Performance of UGS in depleted gas-condensate reservoirs is strongly
composition-dependent. Therefore, it is necessary to predict reservoir fluid compo-
sitional changes under varying pressures and depletion processes. To define the fluid
properties at the reservoir and surface conditions, usually a laboratory PVT analysis,
such as constant volume depletion (CVD) test, is performed on the original reser-
voir fluid. Detailed PVT studies are costly, tedious and time consuming. As a result,
an equation of state (EOS) is often used to match the reported experimental data.
Once matched, the EOS can be used to predict the fluid properties over a wide range
of pressure and temperature conditions. The most successful EOS for natural gas
property calculation is the one proposed by Peng and Robinson. The compositional
analysis of the reservoir fluid is the primary input data to the EOS. Many studies
have emphasized the C7+ characterization as the key element in attaining agreement
between EOS and laboratory results.
Studies show that the composition of injection gas affects condensate recovery
during I/W cycles. Azin et al. defined condensate production enhancement (CPE)
by Eq. (4.1), to determine the role of sour compounds on the condensate recovery in
UGS [20]:

Npso − Npsw
CPE = (4.1)
Npsw

where
Npso : Cumulative condensate production with sour gas injection
Npsw : Cumulative condensate production with sweet gas injection.
Condensate holding ratio (CHR) of injected gas is defined as the ratio of conden-
sate in injected sour gas in equilibrium with reservoir fluid to that in pure CH4 (as a
base component for storage gas) in equilibrium with reservoir fluid (4.2) [20]:
in j.
yC5+
CHR = C H4
(4.2)
yC5+

The CHR can be interpreted as a function of the condensate equilibrium ratio. The
K-value of condensate (C5+ ) in the mixture of injected gas is defined as Eqs. (4.3)
and (4.4) for storage of sour and pure methane, respectively:
160 R. Azin and A. Izadpanahi

sour
yC5+
sour
K C5+ = sour (4.3)
xC5+
pur e
pur e yC5+
K C5+ = pur e (4.4)
xC5+

Also, the mole fraction of condensate in the gas phase for each state is obtained
by rearrangement of Eqs. (4.5) and (4.6).
sour
yC5+ = K C5+
sour sour
xC5+ (4.5)

pur e pur e pur e


yC5+ = K C5+ xC5+ (4.6)

By substituting Eqs. (4.5) and (4.6) in (4.2), the condensate holding ratio is
obtained, as follows:
sour sour
K C5+ xC5+
CHR = pur e pur e (4.7)
K C5+ xC5+

C5+ is the major constituent of the reservoir liquid phase and accounts for more than
85 mol%, regardless of the type of storage injection gas. Therefore, changes in the
mole fraction of C5+ in sour gas injection and pure methane injection can be assumed
to be negligible. By this assumption Eq. (4.7) changes as:
sour
K C5+
CHR = pur e (4.8)
K C5+

Malakooti and Azin conducted simulation on sour and sweet gas injection into
a depleted lean gas condensate reservoir. They investigated the condensate re-
vaporization during the gas storage with different compositions. Figure 4.8 shows
the K-value of the condensate increases in the presence of H2 S and CO2 . Also, it can
be concluded that the effect of H2 S is more significant than CO2 [20].
Figure 4.9 indicates the condensate holding ratio (CHR) for different gas compo-
sitions. The trend of this figure is similar to the previous figure. This figure shows
that pure CO2 can significantly increases the condensate production. In contrast, pure
methane has the highest condensate holding ratio in comparison to other gas streams
[20].
Figure 4.10 shows the cumulative field condensate production during 10 I/W
cycles. As it is obvious from this figure, the sweet gas stream has the lowest conden-
sate production. Also, the condensate production increases by increasing the CO2
and H2 S in the injected gas [20].
Figure 4.11 shows the condensate production enhancement (CPE) versus 10 I/W
cycles. The trend of this figure is similar to previous figure. In this figure, the streams
4 Gas Injection for Underground Gas Storage (UGS) 161

0.04

0.035

0.03

0.025
K Value

0.02

0.015
Sweet Natural Gas
0.01 Pure CO2
Pure CH4
0.8CH4+0.2CO2
0.005 0.8CH4+0.2H2S

0
600 1100 1600 2100 2600 3100
Pressure (psia)

Fig. 4.8 K-value of heavy components for different types of storage gas [20]

4
CH4
3.5 0.8CH4+0.2CO2
Condensate Holding Ratio

0.8CH4+0.2H2S
3
CO2
2.5

1.5

0.5

0
400 900 1400 1900 2400 2900 3400
Pressure (psia)

Fig. 4.9 CHR versus pressure [20]

with 8.3, 2.8 and 0.85% mole percent of CO2 + H2 S have 7, 5 and 3% of condensate
production enhancement, respectively [20].
According to Figs. 4.10 and 4.11 the heavy hydrocarbons have high tendency to
vaporize during the withdrawal stage by increasing the percentage CO2 and H2 S in
the injection gas. The molecules of these gases make the London force with heavy
hydrocarbons. Also, the sulfur atoms and oxygen of H2 S and CO2 are electronegative
and both molecules have a polar structure. Besides, the uneven distribution of molec-
ular electron density of heavy hydrocarbon causes the multi-pole state which helps
to H2 S and CO2 to attract the heavy hydrocarbon. Also, the uneven distribution is
stronger in the presence of H2 S and CO2 . These are the reasons of higher condensate
162 R. Azin and A. Izadpanahi

Cumulative Field Condensate Production


2.40E+06
Sweet Gas Stream
2.20E+06 Sour Gas Stream 1
Sour Gas Stream 2
Sour Gas Stream 3
2.00E+06
(STB)

1.80E+06

1.60E+06

1.40E+06

1.20E+06
1 3 5 7 9
I/W Cycle

Fig. 4.10 Cumulative condensate production for the different type of injected gas [20]

8
Condensate Production Enhancement

H2s+CO2=8.3%
7
H2s+CO2=2.8%
6 H2s+CO2=0.85%

5
(%)

0
1 3 5 7 9
I/W Cycle

Fig. 4.11 CPE caused by different mole fraction of CO2 and H2 S [20]

production and higher CPE by increasing the mole percent of H2 S and CO2 in the
injected gas [20].
4 Gas Injection for Underground Gas Storage (UGS) 163

4.9 Reservoir Material Balance in UGS

4.9.1 Reserve Estimate in UGS

There are three broad categories in the estimation of reserves; that is, volumetric,
material balance and production decline analysis. Selection of appropriate reserve
estimation technique depends on the available information. Generally, the range of
uncertainty for a reserve estimate decreases and confidence level increases when
more information is available and the reserve estimate is validated by more than one
technique.
Volumetric methods involve calculation of reservoir hydrocarbons in place
contained in a rock volume. The key unknown in volumetric reserve estimate may
be rock volume, porosity, or fluid saturation. Material balance methods involve the
analysis of pressure behavior as reservoir fluids are withdrawn. Reserve estimate by
material balance requires sufficient production and pressure data, as well as rock
and fluid properties, characteristics of aquifers, and accurate average reservoir pres-
sure and is more reliable than volumetric method. Production decline methods of
reserves estimation involve the analysis of production behavior as reservoir fluids
are withdrawn. Confident application of decline analysis methods requires suffi-
cient period of stable operating conditions after wells established drainage areas.
Factors affecting production decline behavior include reservoir rock and fluid prop-
erties, transient versus stabilized flow, changes in operating conditions and depletion
mechanism. Reserves may be assigned based on decline analysis when sufficient
production data is available. The decline relationship used in projecting production
should be supported by all available data.
For a gas reservoir with known bulk volume Vb (acre-feet), average porosity φ,
average connate water Sw , gas formation volume factor Bg , the standard volume of
gas in place is calculated by Eq. (4.9):

43560Vb φ(1 − Sw )
G= (4.9)
Bg

The general material balance equation for a gas reservoir is written as follows by
applying the law of conservation of mass to the reservoir and associated production:
 
Cw Swi + C f
G(Bg − Bgi ) + G Bgi P + We = G p Bg + Bw W p (4.10)
1 − Swi

For most gas reservoirs, the gas compressibility term is much greater than the
formation and water compressibility, and the second term on the left-hand side of
Eq. (4.10) becomes negligible:

G(Bg − Bgi ) + We = G p Bg + Bw W p (4.11)


164 R. Azin and A. Izadpanahi

When reservoir pressures are abnormally high, this term is not negligible and
should not be ignored. When there is neither water encroachment into nor water
production from a reservoir of interest, the reservoir is said to be volumetric. For a
volumetric gas reservoir, Eq. (4.11) reduces to:

G(Bg − Bgi ) = G p Bg (4.12)

Using expressions for Bg and Bgi substituting them into Eq. (4.12) results in
Eq. (4.13) at isothermal process, i.e. reservoir temperature:

P Pi Pi
=− Gp + (4.13)
Z Zi G Zi

Because Pi , Z i and G are constant for a given reservoir, Eq. (4.13) suggests that
a plot of PZ versus G p yields a straight line with slope of − ZPi Gi and intercept of
Pi
Zi
. If PZ is set equal to zero, which would represent the production of all the gas
from a reservoir, then the corresponding G p equals OGIP, G. As long as the UGS
reservoir is volumetric with no indication of inventory loss, the material balance
straight line approach describes the behavior of UGS reservoir for I/W cycles. In
reality, pressure-inventory relationship deviates from straight line which is a result
of fluctuation in reservoir pressure at the end of I/W. The I/W end effect caused by
rapid injection and withdrawal is reflected as deviations in pressure-inventory plots.
Flanigan discussed different pressure-inventory scenarios which occur for volumetric
and water drive UGS reservoirs [31]. He described cases with loss of inventory due
to gas leakage through gas cap or as a result of improper casing design, migration to
nearby formations, dissolution into saline aquifer, and etc.
In water-drive reservoirs, the relation between G p and PZ is not linear, and water
influx causes less pressure drop than under volumetric control. In the case of conden-
sate production, the gas equivalent of the condensate in SCF/STB must be added to
G p and two-phase compressibility factor must be placed instead of single phase
compressibility [32].
γ
G E = 133000 (4.14)
Mwo

P
For the volumetric naturally fractured gas reservoir, the familiar Z
equation can
be represented as [33]:
 
P Pi Gp
∗ = 1− (4.15)
Z Z i∗ G
∗ Z
Z = (4.16)
1 − ce (Pi − P)
4 Gas Injection for Underground Gas Storage (UGS) 165

Gerami et al. [33] introduced another equation for total compressibility that has
a term accounting for the effects of fracture porosity and fracture compressibility:

ct = cg + cng (4.17)

cng is related to the compressibility of matrix pore volume, fracture pore volume,
and connate water (non-gas components) compressibility:

cng = ce [1 − cg (Pi − P)] (4.18)

where ce is effective compressibility which is defined as follows:

φ1 (cw Swi + c1 ) + c2 φ2
ce = (4.19)
φ1 (1 − Swi ) + φ2

For the stress-sensitive, naturally fractured reservoirs, fracture porosity φ2 and


fracture compressibility c2 can be calculated at the new reservoir pressure Pi+1 [34]:
  13
φi+1 ki+1 log Pki+1 − log Ph
= = (4.20)
φi ki log Pki − log Ph
−1
c2i+1 = (4.21)
pki+1 ln(Pki+1 /Ph )

where Pk is the stress on the fractures minus the average reservoir pressure, i.e.,
the net stress on the fractures and Ph is an apparent healing pressure which is in
the order of 40,000 psia [34]. In conventional p/z material balance, the extrapolated
straight-line trend of measured shut-in pressures (for gas reservoirs) is used to predict
OGIP.
Since most gas wells do not produce at constant flow rate, a constant pressure
difference between average reservoir conditions and flowing conditions cannot be
assumed. For a variable rate system, the difference between reservoir p/z and flowing
p/z is not constant, but a function of flow rate.
Normalized Rate/Normalized Cumulative approach is an alternative to the flowing
p/z plot which is applicable to both oil and gas reservoirs, and works for constant
or variable rate systems. Flexibility and simplicity are the main advantages of this
analysis. The procedure is similar to the pseudo-steady-state approach, but involves
plotting the inverse of pseudo-steady-state equation, so that a declining trend is
produced:

q 1 2qtca Pi 1
=− + (4.22)
Pp b pss G i (ct μz)i Pp b pss
166 R. Azin and A. Izadpanahi

t
μi cti q(t)
tca = dt (4.23)
q(t) μ(P)ct (P)
0

The normalized pressure integral (NPI) is used to develop a vigorous analytical


method for data drawdown without any noise, as is typical of standard well test
derivative. The solution involves a pressure integral curve as the base curve for noisy
drawdown analysis. The dimensionless pressure integral is defined as follows:

t D A
1
PDi = PD (t)dt (4.24)
tD A
0

khPp
PD = (4.25)
1.417 × 106 T q
0.00634ktca
tD A = (4.26)
π φμi cti re2

Conceptually, pressure integral is a cumulative average flowing pressure drop.


The smoothness of pressure integral makes it as an ideal base curve for the standard
pressure derivative, if the raw data contains any degree of noise or scatter.

4.9.2 Drainage Radius in UGS

The drainage radius in gas storage wells is described as the distance from the flow
boundary in the reservoir to the center of production well. Ligen et al. represented a
mathematical relation for calculating gas drainage radius of UGS. Their model was
developed for gas production based on the following assumptions:
• The reservoir is assumed as a cylindrical shape
• A production well is located at the center of the cylinder
• The vertical thickness of the reservoir is sufficient
• The reservoir is homogeneous and isotropic and also has good connectivity
• Gas flow considered as single-phase fluid flow
• Gas produces at a constant rate.
They presented the following equation for calculating the drainage radius in gas
storage wells. More information about the solution techniques and deriving equations
are available in Ligen et al. [35].


Pp max − Ppw f p Bg Tsc ηt p Re p 3


ηt
14.682 2p
−2 = ln − − 0.84e Re p
(4.27)
Psc (Pmax − Pmin )T Re2p rw 4
4 Gas Injection for Underground Gas Storage (UGS) 167

Table. 4.4 The accuracy of Eq. 4.27 reported by Ligen et al. [35]
Well number Permeability, mD Cal drainage radius, m Actual radius, m Relative error, %
1 4 151 169 −11.9
2 9 220 197 10.5
3 13 245 266 −8.6
4 16 289 312 −8
5 25 356 328 7.9
6 30 365 402 −10.1
7 36 422 437 −3.6
8 42 427 447 −4.7
9 49 488 463 5.1
10 55 484 513 −6.0
11 64 553 583 −5.4
12 70 542 577 −6.5
13 81 618 634 −2.6
14 86 598 626 −4.7
15 92 617 669 −8.4
16 96 629 674 −7.2
17 100 683 672 2.6

where
K
η=
μφct

Table 4.4 indicates the Ligen et al. study of drainage radius for 17 UGS wells.
Their results showed that the radius values calculated by the mathematical model is
accurately matched to the field data. Also, they expressed that the drainage radius
depends on multiple variables. By increasing the rock permeability and production
time, the drainage radius increases [35].

4.10 Well Pattern in UGS

Proper design of wells is an important step in the development of UGS. Well number,
well location, well completion, and well type, i.e. horizontal or vertical are key
factors in an optimum well pattern study. Application of horizontal drilling in UGS
is emerging in a few fields. In 2000, about 100 horizontal wells were operating out of
some 10,000 storage wells in 600 UGS facilities in the world [36]. Simultaneously,
at least more than 100 new horizontal wells were planned, pointing out the growth
in the use of this technique to increase maximum deliverability. Studies show that
168 R. Azin and A. Izadpanahi

a horizontal well intercepting vertical fractures has a better productivity index than
a vertical well intercepting a vertical fracture. Horizontal wells have productivity
index 1.5–6 times vertical well in the same reservoir, depending on reservoir quality
and horizontal drain length [36]. Horizontal wells can also minimize water conning
during operation, if the drain trajectory is always above the gas–water interface and
pressure drop inside the drains is smaller than in a vertical well, so less water coning
during withdrawal [36]. Tureyen et al. [5] investigated the effects of well properties
on the performance of UGS reservoir in Northern Marmara gas field in Turkey. Their
modeling studies showed that the wellhead pressure has significant effect on the
performance of UGS reservoir. For a fixed number of wells, working gas capacity
increases as the wellhead pressure is decreased. They found that the number of
horizontal wells required to obtain the desired working gas capacity strongly depends
on the wellbore damage, horizontal lengths and partial completion characteristics.
Efforts to minimize the wellbore damage help to improve the production/injection
performance of the storage reservoirs.
Recently, application of horizontal well is commenced in the development of
UGS in LiaoHe oil field, China [37]. The authors address technical challenges for
converting UGS in complex fault-block reservoir, including local environmental
restrictions and imperfect gas storage well cementing technology in LiaoHe oil
region. Under these circumstances, horizontal wells and branching wells drilling
and completion technologies are suggested to complete gas storage drilling.
For instance, a working gas capacity of 1 billion sm3 is estimated to be obtained
by using 10 horizontal wells each with a horizontal length of 100 m if there is no
wellbore damage. Keeping the well number the same, 1 billion sm3 is obtained by
increasing the horizontal length nearly five times if the mechanical skin factor is 20.
The advantages of using horizontal wells include intercepting natural fractures,
higher recoveries, reduction of water conning and reduction of sand production prob-
lems. On the other side, horizontal wells are costly compared to vertical wells. In a
UGS plan, new wells may be drilled near the crests and in high gas saturation and
permeability zones, especially in the high permeability fractured zones. Having high
permeability leads to increase connection transmissibility factor; this in turn results
in well productivity or injectivity improvement. Productivity index is defined as the
ratio of production rate of a given phase to pressure drawdown defined as follows:

Qp
J= (4.31)
Pr − Pw

where Q p indicates the production rate of a given phase and Pr presents the pressure
in the reservoir at the drainage radius. Equation (4.31) can be applied for injection
wells if injection rates are treated as negative.
4 Gas Injection for Underground Gas Storage (UGS) 169

4.11 Monitoring of UGS Projects

4.11.1 Surface Monitoring

Surface monitoring techniques are divided into conventional monitoring and


synthetic aperture radar (SAR). Conventional techniques include geotechnical moni-
toring (clinometers, distometers, extensometers and etc.), GPS (global positioning
system) or conventional topographic survey. These conventional techniques can be
applied to limited set of benchmarks and also it needs the access to the area. SAR
technology has high potential for evaluating ground surface displacement over wide
areas. This technology collects a vast number of data from remote sensing satellites
which prepare a deformation map. This technology doesn’t need any physical contact
and ground access for preparing the deformation map of the monitoring area [19].

4.11.2 Well Monitoring

Static bottom-hole pressure is one of the characteristics that measures at the end
of each injection cycle and at the end of each withdrawal cycle. This characteristic
evaluates the reservoir connectivity which specifies the performance of the UGS field
and gas inventory. Saturation variation is recorded to verify the place of gas water
contact using RST logs. Also, the well head pressure and gas rates are monitored
using telemetry for well performance evolution. Mini-frac test or leak-off test is
performed for ensuring the mechanical integrity of the seal rock [19].

4.11.3 Wellbore Integrity Monitoring

A most important requirements for UGS wells is the efficient zonal isolation which
catered by the cement between casing and wellbore. The zonal isolation may loss
due to high injection pressure and production rates. Also, pressure variations may
induce micro-annulus or micro cracks in the cement over time. Cement failure can
caused the loss of efficient zonal isolation, gas migration behind the casing and
corrosion problems. Cement distribution can be monitored by wireline logging with
ultra-sonic and sonic tools. Also, these tools can identify the discrete defects such as
micro-annulus and channeling [19].
Azin et al. presented a new formulation for the cement used in UGS wells.
The advantages of this formulation include improvement of cement ductility and
elasticity, preventing cement cracking and decreasing the crack propagation. A
polymer and an elastomer will be added to the common cement composition. Their
results showed that by adding the latex to the cement formulation, the hydration,
density, fluid loss and bonding of the cement are decreased. Also, they expressed
170 R. Azin and A. Izadpanahi

that polypropylene fiber can significantly improve the stress–strain properties and
compressive strength of the cement. On the other hand, the combined flexibility of
microfibers and cohesive effect of latex can increase the resistance of the cement
against excessive pressures [38].

4.11.4 Micro-Seismic Monitoring

Pressure variation and gas movements can be tracked with micro-seismicity method.
In this method, micro-seismic events are detected by passive seismic monitoring.
This detection is done by the creation or reactivation of small fractures or by the
modification of pore pressures. This method is provided a real-time picture of the
passage of gas at specific points where the appearance of micro-seismicity and the
increase in pore pressure are linked together. Also, the weakness zones in the reservoir
or cap rocks can be characterized by this method [19].

4.12 Numerical Simulation and Dynamic Modelling of UGS

The objective of reservoir simulation in UGS is to predict reservoir performance, pres-


sure distribution, storage capacity, productivity index, and etc. In general, numerical
simulation is used to evaluate the potential of depleted reservoirs for switch to gas
storage reservoirs. Parameters affecting UGS simulation include reservoir geometry,
rock properties, PVT of injecting and reservoir fluid, well pattern and well placement.
Optimum base pressure and optimum time to start UGS, gas volume remaining in
reservoir at the start of UGS, I/W scenarios, optimum number, location and type of
wells, and etc. are determined by numerical reservoir simulation. Details of governing
equation, discretizing the partial differential equations, and solution techniques in
the context of reservoir simulation are discussed by Ertekin et al. [39]. Some issues
must be considered in the dynamic modelling of UGS operations, including gas
composition variation, reservoir temperature variation due to injection of cold gas,
pressure variation due to the repeated I/W cycles at high gas rates, hysteresis of
capillary pressures, rock compressibility and gas water relative permeabilities. The
mentioned issues are not considered in the common reservoir simulation. Another
difference between UGS and conventional reservoir simulation is the time steps. In
the UGS simulation, the frequent switches between injection and withdrawal should
be correctly accounted for. But, in the conventional reservoir simulation, the time
steps are typically set (yearly, monthly and etc.) to optimize the simulation run time
[19].
Xiao et al. [4] carried out a simulation study using both 3D full-field black oil reser-
voir model and full-field multi-component reservoir model to evaluate and consider
the feasibility of creating UGS plan in a buried hill fractured oil depleted carbonate
reservoir with bottom water, sufficient closure, a good quality and tight cap rock.
4 Gas Injection for Underground Gas Storage (UGS) 171

Khamehchi and Rashidi [14] studied simulation of UGS in Sarajeh gas field, Iran.
Based on available information and those obtained by the model, they concluded
this field could be a candidate for UGS plan. Later, this reservoir was converted
in to the first operating UGS reservoir in Iran. Compositional model was used to
create the model in their studies. Azin et al. [21] simulated underground gas storage
in a partially depleted gas reservoir using a compositional simulator. The purpose
of their research was optimization of performance of UGS with different reservoir
depletion scenarios, gas injection and aquifer strength. They concluded that if the
gas reservoir is depleted down to its ultimate recovery, it may contain less base gas
than required by gas storage operations. Under these conditions, the withdrawal rate
from reservoir may not meet the target rate in the high-consumption seasons. At
these conditions, operator may set the withdrawal at a lower target rate and let the
reservoir take enough pressure before raising the target pressure. Alternatively, it can
prolong the injection period in the first cycle in order to inject sufficient reserve into
reservoir before starting withdrawal. In separate studies, effect of sour gas injection
as a strategic UGS, feasibility of UGS in a fractured reservoir, and optimization of
well pattern and well location was presented by Azin et al. [20], Jodeyri Entezari
et al. [22], and Malakooti and Azin [8]. Soroush and Alizadeh [13] presented a real
case study of an Iranian gas condensate reservoir for UGS. The feasibility of doing
a storage project was supported by the simulator results, and the reservoir is capable
of storing the gas and withdrawing it in wintertime without any problem. Based on
the simulation results, they concluded that it is possible, and even suitable, to main-
tain gas storage with higher field pressures. McVey and Spivey [40] used computer
reservoir simulation to maximize the deliverability of the gas storage reservoir and
minimize the cost to meet a particular I/W schedule.
Presence of naturally occurring fractures incorporates additional complexity to
the behavior of UGS reservoir. Gerov and Lyomov [41] analyzed the performance
of UGS in a multi-layered, fractured carbonate and sandstone formation. Simula-
tion confirmed that water encroachment through fractures and water cut result in the
unstable UGS performance and gas dis-balance. Bilgesu and Ali [42] presented a
numerical model to predict the impacts of important formation properties such as
matrix permeability, fracture spacing, and fracture orientation on the efficiency of
well design for UGS in a fractured reservoir. Jodeyri Entezari et al. [22] investigated
UGS in a partially depleted, naturally fractured gas reservoir through compositional
simulation. Effects of fracture parameters, i.e. fracture shape factor, fracture perme-
ability and porosity were studied. They concluded that water production from reser-
voir is affected by the distribution of fracture density. Also, fracture shape factor
affects water flow through porous medium, and active aquifer can reduce condensate
drop out around the well bore. For a complete UGS case study, the reader is referred
to Malakooti and Azin which the results of this study were explained in Sect. 4.8 [3].
172 R. Azin and A. Izadpanahi

References

1. Bahmannia G. Developing gas markets in persian gulf, case study: Iran. In: 23rd gas conference.
Amesterdam;s 2006.
2. Tek MR. Natural gas underground storage: inventory and deliverability. Pennwell Books; 1996.
3. Malakooti R, Azin R. Simulation study of underground gas storage. Germany: Lambert
Academic Publishing; 2012.
4. Xiao G, Du Z, Ping G, Du Y, Yu F, Tao L. Design and demonstration of creating underground
gas storage in a fractured oil depleted carbonate reservoir (Russian). In: SPE Russian oil and
gas technical conference exhibition. Society of Petroleum Engineers; 2006.
5. Tureyen OI, Karaalioglu H, Satman A. Effect of the wellbore conditions on the performance
of underground gas-storage reservoirs. In: SPE/CERI gas technology symposium. Society of
Petroleum Engineers; 2000.
6. Cornot-Gandolphe S. Underground gas storage in the world—2017 status. Cedigaz Insights;
2017.
7. Bennion DB, Thomas FB, Ma T, Imer D. Detailed protocol for the screening and selection
of gas storage reservoirs. In: SPE/CERI gas technology symposium. Society of Petroleum
Engineers; 2000.
8. Malakooti R, Azin R. The optimization of underground gas storage in a partially depleted gas
reservoir. Pet Sci Technol. 2011;29:824–36.
9. Coffin P, Lebas G. Converting the pecorade oil field into an underground gas storage. SPE Proj
Facil Constr. 2008;3:1–6.
10. Anyadiegwu CIC. Development of depleted oil reservoirs for simultaneous gas injection for
underground natural gas storage and enhanced oil recovery in Nigeria. In: SPE Nigeria annual
international conference exhibition. Society of Petroleum Engineers; 2016.
11. Mgbaja UM, Enwere N. Reservoir characterization, simulation & estimation of storage capacity
of depleted reservoirs in Niger Delta for Underground Natural Gas Storage. In: SPE Nigeria
annual international conference exhibition. Society of Petroleum Engineers; 2017.
12. Burke WF. Simultaneous underground gas storage and secondary oil recovery. J Pet Technol.
1960;12:22–6.
13. Soroush M, Alizadeh N. Underground gas storage in partially depleted gas reservoir. In:
Canadian International Petroleum Conference. Petroleum Society of Canada; 2007.
14. Khamehchi E, Rashidi F. Simulation of underground natural gas storage in Sarajeh Gas Field,
Iran. In: SPE Technical Symposium of Saudi Arabia Section. Society of Petroleum Engineers;
2006, p. 1–14. https://doi.org/10.2118/106341-MS.
15. Yanze Y, de Kok JH, Clemens T. Optimised combined underground gas storage and enhanced
oil recovery. In: European conference and exhibition. Society of Petroleum Engineers; 2009.
16. Yang C, Wang T, Li J, Ma H, Shi X, Daemen JJK. Feasibility analysis of using closely spaced
caverns in bedded rock salt for underground gas storage: a case study. Environ Earth Sci.
2016;75:1138.
17. Ojukwu C, Smith K, Kadkhodayan N, Leung M, Dame R, Voskanian A. Case study: natural
gas storage in federal waters offshore California. In: SPE western regional meeting. Society of
Petroleum Engineers; 2019.
18. Tien C, Lin H, Chen J, Wu W. Case study of underground gas storage in a lean gas condensate
reservoir with strong water-drive. In: SPE/IATMI Asia Pacific oil gas conference exhibition.
Society of Petroleum Engineers; 2019.
19. Verga F. What’s conventional and what’s special in a reservoir study for underground gas
storage. Energies. 2018;11:1245.
20. Azin R, Malakooti R, Helalizadeh A, Zirrahi M. Investigation of underground sour gas storage
in a depleted gas reservoir. Oil Gas Sci Technol d’IFP Energies Nouv. 2014;69:1227–36.
21. Azin R, Nasiri A, Entezari J. Underground gas storage in a partially depleted gas reservoir. Oil
Gas Sci Technol l’IFP. 2008;63:691–703.
22. Jodeyri Entezari A, Azin R, Nasiri A, Bahrami H. Investigation of underground gas storage in
a partially depleted naturally fractured gas reservoir. Iran J Chem Chem Eng. 2010;29:103–10.
4 Gas Injection for Underground Gas Storage (UGS) 173

23. Zirrahi M, Azin R, Hassanzadeh H, Moshfeghian M. Prediction of water content of sour and
acid gases. Fluid Phase Equilib. 2010;299:171–9.
24. Zirrahi M, Azin R, Hassanzadeh H, Moshfeghian M. Mutual solubility of CH4, CO2, H2S, and
their mixtures in brine under subsurface disposal conditions. Fluid Phase Equilib. 2012;324:80–
93.
25. Azin R, Nasiri A, Entezari AJ, Montazeri GH. Investigation of underground gas storage
in a partially depleted gas reservoir. In: CIPC/SPE gas technology symposium 2008 Joint
Conference. 2008. p. 18. https://doi.org/10.2118/113588-MS.
26. Adel H, Tiab D, Zhu T. Effect of gas recycling on the enhancement of condensate recovery,
case study: Hassi R’Mel South field, Algeria. In: International oil conference exhibition in
Mexico. Society of Petroleum Engineers; 2006.
27. Zaitsev IY, Dmitrievsky SA, Norvik H, Yufin PA, Bolotnik DN, Sarkisov GG, et al. Compo-
sitional modeling and PVT analysis of pressure maintenance effect in gas condensate field:
comparative study. In: European petroleum conference. Society of Petroleum Engineers; 1996.
https://doi.org/10.2118/36923-MS.
28. Seto CJ, Jessen K, Orr Jr FM. Compositional streamline simulation of field scale condensate
vaporization by gas injection. In: SPE reservoir simulation symposium. Society of Petroleum
Engineers; 2003.
29. Melean Y, Bureau N, Broseta D. Interfacial effects in gas-condensate recovery and gas injection
processes. In: SPE annual technical conference exhibition. Society of Petroleum Engineers;
2001.
30. Wichert E, Aziz K. Calculate Zs for sour gases. Hydrocarb Process. 1972;51:119.
31. Flanigan O. Underground gas storage facilities: design and implementation. Elsevier; 1995.
32. Terry RE, Rogers JB, Craft BC. Applied petroleum reservoir engineering. Pearson Education;
2015.
33. Gerami S, Pooladi-Darvish M, Mattar L. Decline curve analysis for naturally fractured gas
reservoirs: a study on the applicability of “pseudo-time” and “material balance pseudo-
time”. In: International petroleum technology conference. International Petroleum Technology
Conference; 2007.
34. Jones FO Jr. A laboratory study of the effects of confining pressure on fracture flow and storage
capacity in carbonate rocks. J Pet Technol. 1975;27:21–7.
35. Ligen T, Guosheng D, Chunhui S, Jieming W. Mathematical description of gas drainage radius
for underground gas storage. Chem Technol Fuels Oils. 2018;54:500–8.
36. Bagci AS, Ozturk B. Performance analysis of horizontal wells for underground gas storage in
depleted gas fields. In: Eastern regional meeting. Society of Petroleum Engineers; 2007.
37. Zhu Z, Yang Y. Horizontal drilling and cementing technology of the gas storage in LiaoHe
oil region. In: IADC/SPE Asia Pacific drilling technology conference exhibition. Society of
Petroleum Engineers; 2012.
38. Shahvali A, Azin R, Zamani A. Cement design for underground gas storage well completion.
J Nat Gas Sci Eng. 2014;18:149–54.
39. Ertekin T, Abou-Kassem JH, King GR. Basic applied reservoir simulation. Reservoir: Society
of Petroleum Engineers; 2001.
40. McVay DA, Spivey JP. Optimizing gas storage reservoir performance. In: SPE annual technical
conference exhibition. Society of Petroleum Engineers; 1994.
41. Gerov LG, Lyomov SK. Gas storage performance in a fractured formation. In: SPE gas
technology symposium. Society of Petroleum Engineers; 2002.
42. Bilgesu HI, Ali W. Effect of reservoir properties on the performance and design of gas storage
wells. In: SPE eastern regional meeting. Society of Petroleum Engineers; 2004.
Chapter 5
Gas Injection for Pressure Maintenance
in Fractured Reservoirs
Ahmad Jamili, Amin Izadpanahi, Pooya Aghaee Shabankareh,
and Reza Azin

Abstract Gas injection into the gas cap which is known as pressure maintenance
or crestal gas injection is done to increase the reservoir pressure. Different types
of gas may be injected in this method including producing gas, N2 , CO2 etc. The
injected gas is chosen base on the field development studies. Each of these gases has
some advantage and disadvantages. Gas injection in naturally fractured reservoirs is
a challenge which needs more investigation on this subject. This chapter summarizes
the basic concepts of pressure maintenance and active mechanisms during pressure
maintenance in naturally fractured reservoirs. Also, this chapter provides the essential
concepts in simulation of pressure maintenance in fractured reservoirs.

Abbreviations

Acronyms

AD-OO Automatic differentiation-object oriented


BHP Bottom-hole pressure
EC European commission
EDFM Embedded discrete fracture model
EOR Enhanced oil recovery

A. Jamili
Saint Francis University, Loretto, PA, USA
e-mail: ajamili@francis.edu
A. Izadpanahi · P. Aghaee Shabankareh
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 175
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_5
176 A. Jamili et al.

GMRES Generalized minimal residual method


GOR Gas oil Ratio
HFM Hierarchical fracture module
ILU Incomplete lower/upper
IOR Improved oil recovery
IPR Inflow-performance relation
MINC Multiple Interacting Continua
MRST Matlab reservoir simulation toolbox
MsRSB Multiscale restriction smoothed basis
SAIGUP Sensitivity analysis of the impact of geological uncertainties on
production forecasting

Variables

A Area, m2
B Formation volume factor, vol/vol
C Concentration, mole/m3
c Component
cf Formation Compressibility, 1/psi
D Depth, m
Dc,o , Dc.g Diffusion coefficient of component c in oil and gas, cm2 /s
Di j Binary diffusion coefficient of components i and j, cm2 /s
Dg Gas diffusion coefficient, cm2 /s
De,c Effective diffusion coefficient for component c at matrix-fracture
boundary, cm2 /s
De,i Effective diffusion coefficient for component i, cm2 /s
d Correlation coefficients
e Correlation coefficients
f o,c , f g,c Fugacity of component c in oil and gas, psi
f m,i Fugacity of component i in phase m, psi
F Formation resistivity factor
G Gas in place, ft3
H Fracture thickness in z-direction, m
k Permeability, md
kc Diffusion mass transfer coefficient of component c at matrix-fracture
boundary, mole/(m2 s)
Kc Equilibrium ratio of component c
ki,j Binary interaction coefficient
kro , krg ,krw Relative permeability of oil, gas, and water
krgcw Gas relative permeability at connate water
krocw Oil relative permeability at connate water
krwro Water relative permeability at residual oil saturation
L Moles of oil per unit mole feed
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 177

l Length of fracture, m
m Cementation factor
MWi Molecular weight of component i, g/gmole
n1 Exponent
nc Number of components
N Oil in place, bbl
Nc,p Diffusion molar flux of component c at phase p, mole/(m2 s)
nog ,ng ,nw ,now Exponents on relative permeability curves
Pcog , Pcow Capillary pressure (oil–gas and oil–water), psi
Pc0 Reference capillary pressure at reference interfacial tension, psi
p Pressure, psi
Pc Capillary pressure, psi
pc,i Critical pressure of component i, psi
Pi Parachor of component i
po ,pg ,pw Pressure of oil, gas, and water, psi
pr e f Reference pressure, psi
p Pressure gradient, psi/ft
q D, f m,c Diffusion rate of component c at the matrix-fracture boundary,
mole/s
qC, f m,c Convection mass transfer rate of component c at the matrix-fracture
boundary, mole/s
q Flow rate, ft3 /day
R Universal gas constant, cm3 MPa/(K. mole)
Rs Solution gas oil-ratio, scf/stb
Rs Produced gas oil-ratio, scf/stb
Sgg Geometric mean of matrix and fracture gas saturation
So ,Sg ,Sw Saturation of oil, gas, and water
Sgr Residual gas saturation
Sorg Residual oil saturation to gas
Sorw Residual oil saturation to water
Swc Critical water saturation
Swir Irreducible water saturation
Si Volume shift parameter in PR EOS
t Time, day
T Temperature, K
Tc,i Critical temperature of component i, K
To ,Tg ,Tw Transmisibilities of oil, gas, and water, mole.md/(m2 .cp)
M
To,c , Tg,cM
Molecular transmisibilities of component c in oil and gas, mole/s
Tr,i Reduced temperature of component i
V Moles of vapor per unit mole feed
Vp Pore volume
v Average gas stream velocity in the fracture, m/s
vo Oil bulk velocity, m/s
vg Gas bulk velocity, m/s
v x , v y , vz Fluid bulk velocities in x, y, and z directions, m/s
178 A. Jamili et al.

Vr Bulk volume, m3
Vp Pore volume, m3
Vc,i Critical volume of component i, cm3 / mole
W Fracture width in y-direction, m
x,y,z Cartesian coordinates
xc Mole fraction of component c in oil phase
xj Mole fraction of component j in oil phase
xi,m ,xj,m Mole fraction of component i and j in phase m
yc Mole fraction of component c in gas phase
yj Mole fraction of component j in gas phase
yc,m f Mole fraction of component c in the gas phase at matrix-fracture
boundary
yc, f Mole fraction of component c at the entrance of the fracture
(yi )m , (yi ) f Mole fraction of component i in gas phase in matrix and fracture
Zc Overall composition of component c
Zj Overall composition of component j
Zo ,Zg ,Zm Compressibility factor of oil, gas, and phase m
zr e f Reference elevation, m
We Water influx

Greek

αs Factor for considering skin-effect at matrix-fracture boundary


i j Collision diameter of the Lennard–Jones potential
σi j Collision integral of the Lennard–Jones potential
t Time step, day
x, y, z Grid cells dimensions, m
γo , γ g , γw Specific gravity of oil, gas, and water, psi/ft
γo , γ g , γw Average specific weight of oil, gas, and water, psi/ft
μo , μg , μw Viscosity of oil, gas, and water, cp
φ Porosity
φ0 Porosity at a reference pressure
φo,c , φg,c Fugacity coefficient of component c in oil and gas
ρo , ρg , ρw Molar densities of oil, gas, and water, mole/cm3
ρr Reduced density
ρm Mixture molar density, mole/cm3
ρC,s Critical density of component c, mole/cm3
ρmr Reduced density of the mixture
λ Mobility
σ Interfacial tension, dyne/cm
σ0 Initial interfacial tension corresponding to the read-in capillary
pressure, dyne/cm
τ Tortuosity of the porous medium
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 179

ωi , ωc Acentric factor of component i and c

Subscripts

c Component
c Capillary
c Critical
f Fracture
g Gas
i Component
i Grid block number in x-direction
i Initial
j Grid block number in y-direction
k Grid block number in z-direction
m Mixture
m Phase
m Matrix block
o Oil
p Phase
p Produced
p Pore
r Reduced
ref Reference
t Total
x,y,z X,y,z directions
w Water

Superscript

l Iteration level
L Time step
M Molecular diffusion
v Vapor
180 A. Jamili et al.

5.1 Introduction

Gas injection into the gas cap which is known as pressure maintenance or crestal
gas injection is done to increase the reservoir pressure. Different types of gas may
be injected in this method including producing gas, N2 , CO2 etc. The injected gas is
chosen base on the field development studies. Each of these gases has some advantage
and disadvantages. Gas injection in naturally fractured reservoirs is a challenge which
needs more investigation on this subject. This chapter summarizes the basic concepts
of pressure maintenance and active mechanisms during pressure maintenance in
naturally fractured reservoirs. Also, this chapter provides the essential concepts in
simulation of pressure maintenance in fractured reservoirs.
It is common to inject gas in naturally fractured reservoirs to maintain the reservoir
pressure and increases oil recovery primarily by gravity drainage and to a lesser extent
by mass transfer between the flowing gas in the fracture and the porous matrix.
In most cases, mass transfer is considered to contribute a small amount to the oil
displacement. Mass transfer could be an important recovery mechanism in the case
of a low permeable and/or small matrix block size. The mechanism is aided by the
high area is available for mass transfer in naturally fractured reservoirs. Although
gravity drainage has been studied extensively, there has been limited research on
mass-transfer mechanisms between the gas flowing in the fracture and fluids in the
porous matrix.
Some mathematical models were developed which describes the mass transfer
in between a gas flowing in the fracture and resident fluid in a matrix block. The
injected gas diffuses into the porous matrix through gas and liquid phases causing
the vaporization of oil in the porous matrix which is transported by convection and
diffusion to the gas flowing in the fracture. Mass transfer between the fracture and
the matrix is assumed to occur by diffusion mass transfer and fluid flow between the
matrix and the fracture. The model accounts for diffusion and convection mechanisms
in both gas and oil phases in the porous matrix driven by capillary pressure gradients
which are generated due to changes in phase behavior as the gas dissolves in the oil
phase.

5.2 The Concept of Pressure Maintenance

Hydrocarbons can be produced using natural reservoir energy. This stage of produc-
tion is called natural depletion or primary production. In this stage, production accom-
plished under various schemes such as gravity drainage, solution gas drive, rock and
fluid expansion, gas cap expansion or aquifer drive. The dominant primary production
mechanism depends on the reservoir properties and the production regime.
The reservoir pressure and energy decrease due to the fluid production. At this
stage, it is necessary to augment the natural energy with an external source. IOR and
EOR methods are used as the external sources. IOR methods accomplished by water
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 181

and/or gas injection. The main purpose of IOR is to maintain reservoir’s pressure.
Hence, IOR methods are usually called pressure maintenance [1]. In the case of
strong reservoir aquifer, it is necessary to maintain gas cap pressure equal to aquifer
pressure to retard the water encroaching the reservoir and controlling water–oil ratio.
Gravity difference between oil and injected gas plays an important role through the
oil production. Gravity force determines fluids distribution and through a balance
with capillary force leads to oil production [2].
Oil production can be high from a reservoir in which pressure is maintained by gas
injection and still have an active gravity segregation. The velocity of gas flow must
be controlled so that it does not dissolve in the oil but push the oil downward. This
velocity is called critical velocity of counterflow which determines oil extraction.
Counterflow velocity highly depends on the rate of pressure drop. It is better to
maintain reservoir’s pressure above solution pressure where no gas would evolve
and there would be no counterflow of gas and oil.
In gas-oil gravity process, gas is injected in the gas cap through vertical wells and
oil is produced by the horizontal wells at the bottom of oil zone. Gas fills the voidage
which was created by oil production and overcomes pressure drop in both oil and
gas zone. Injected gas also recompresses dispersed gas into oil zone which leads to
gravity drainage process [3].
As the rate of injection or production decreases, recovery by this scheme increases.
Capillary force act opposes of gravity force, so as the pores are depleted, gravity
drainage rate will be slower [4, 5]. If there is a considerable difference between
mobility ratio of gas and oil, the counter-flow rate is largely governed by the smaller
mobility ratio [6].

5.2.1 Gas Injection in Gas-Cap

Pressure maintenance is a process which either all the reservoir’s gas or part of
it or an internal gas will be injected into the gas-cap at any stage of production
to supply production energy and overcome the pressure drop [7]. Pressure mainte-
nance partakes unused gas to maintain pressure and keep natural depletion mech-
anisms process [8]. The best recovery is obtained in reservoirs with high vertical
permeability, a thick column of oil or in fractured reservoirs [9].
Natural gases in oil reservoirs may be present in free or solution condition. In the
case of existing free gas, the reservoir may have primary gas cap. But, in the case of
no free gas, the solution gas is evolved from the oil by decreasing the pressure below
the bubble point pressure and secondary gas cap is formed [10, 11].
Solution gas drive starts under bubble point pressure, where gas evolves from oil.
This evolved gas causes oil to flow in the direction of pressure declines [5]. The
evolved gas is also helped the gravity drainage mechanism for producing more oil
[12].
Craig and Geffen [13] compared oil recovery from sandstone reservoirs from five
producing programs as (a) solution drive, (b) full pressure maintenance at original
182 A. Jamili et al.

Cumulative Surplus Gas at Atomospheric 180


Solution Drive
160 PPM (pi=2000 psig)
140 PPM (pi=2370 psig)
Conditions (pore volumes)

FPM (pi=2000 psig)


120 FPM (pi=2370 psig)

100

80

60

40

20

-20
0 5 10 15 20 25 30
Oil Recovery (%)

Fig. 5.1 Oil recovery under partial pressure maintenance (PPM) and full pressure maintenance
(FPM) [13]

bubble point (2370 psi), (c) partially pressure maintenance at 2370 psi, (d) starting
full pressure maintenance after depletion to 2000 psi, (e) starting partially pres-
sure maintenance after depletion to 2000 psi. Figure 5.1 shows the recovery under
different schemes which indicates that best recovery will be obtained by full pressure
maintenance at original bubble point with 10,000 cu ft/bbl GOR.
The gas injection for pressure maintenance is not recommended in some reservoirs
especially in reservoirs with high producing gas-oil ratio. Izadpanahi et al. studied
different scenarios for improving oil recovery in an Iranian fractured oil reservoir.
Their results showed that by increasing the rate of gas injection in the studied reservoir
the final recovery decreased [14].

5.2.2 Gas Recycling

Produced gas is reinjected into the reservoir in the gas recycling projects [15, 16].
In gas condensate reservoirs, when the reservoir pressure decreases to below the
dew point pressure, light components (such as methane) leave the solution and a
heavy liquid phase called condensate form [17]. The maximum pressure drop occurs
near production wells, so the maximum amount of condensate is expected to form
in this region [18]. The main purpose of gas recycling is to prevent the loss of
retrograde liquid phase formed in the reservoir. Increasing pressure above dew point
can postpone the formation of condensate [19–21]. Also, the produced gas in the oil
reservoirs is reinjected with the purposes of pressure maintenance. More information
about gas recycling will be provided in Chap. 6.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 183

5.2.3 Water Injection

Water is injected into the reservoir to maintain pressure and to sweep the oil. Water
injection is known as a method of secondary oil recovery. This method has been
widely used around the world, San Jorge Basin [22], Natih field [23], Zarzaitine field,
Algeria [18], Mamou field [17], the Fort Chadbourne field [24], Adena sand field
[25], Greater Burgan field [26], Carbonate oil field, Abu Dhabi [27], wainwright and
Wildmere fields, Canada [28, 29], the Elk Hills oil field, California [30], Southern
Louisiana gas reservoir [31], Barmer basin of Rajasthan, India [32], Wilmington
field, Los Angles [33], the Battrum Northeast pool, Saskatchewan [34], Inglewood
field, California [35], Siri oil field, Iran [36] and many others [37–44].

5.3 History of Pressure Maintenance

Irwin L. Dunn was the first one to demonstrate the benefits of gas injection into
a shallow producing well in 1903. He believed that pressure maintenance retains
reservoir energy. Gas and/or air was injected in Pennsylvania and West Virginia
before 1910 and, by 1911 pressure maintenance was started in Ohio. Significant
increasing in oil production was observed in Ohio which makes pressure maintenance
a proven method for increasing recovery [8].
In the early days of pressure maintenance, injection was performed at low pressure
of below 100 psi using vacuum pumps. Later it becomes obvious that using vacuum
pumps does not increase ultimate recovery appropriately. Lease operation limitation
was the main problem in the early days of repressuring. The first operation with a large
portion of a field as a unit was in Nowata County, Oklahoma. Finally the first case
of gas injection with unitization in the usual sense was in Kettleman Hills reservoir
in California [4]. The first pressure maintenance by gas injection was started in the
1920s in the Macksburg field, Washington improved techniques and theories [15]. In
1980’s in Venango County, Pennsylvania, an operator, James Dinsmoor, purposely
allowed communication between a partially depleted oil sand and an underlying gas
sand. A few years, he installed a gas pump for injecting gas into depleted wells [4].

5.4 Sources of Injected Gas

Produced gas. Produced gas is easy to get and needs no additional facilities for
capturing. Reinjecting reservoir gas conserves reservoir energy. The economic feasi-
bility of natural gas injection is doubted in some area due to the price and usage
of natural gas in that area. Today, gas plays an important role in the industry. As
mentioned earlier, the produced gas is injected into the reservoir to maintain reser-
voir pressure and increase final recovery of hydrocarbons. Produced gas injection in
184 A. Jamili et al.

the Shoats Creek field recovered 50 percent of oil in place [45]. The Pickton field
Bacon Lime reservoir [46] was under pressure maintenance by gas and water injec-
tion. Only 6.6 MSTB oil in place of total 34 MMSTB can be produced by primary
production. After gas injection, GOR increased which caused shut-in of 100 wells of
total 166 wells. By controlling high GOR, final recovery from this reservoir reached
to 25 MMSTB.
In another case, all of the produced gas from Ekofisk reservoir was reinjected
which increased recovery by 18% of oil in place [47]. Reinjecting produced gas
needs a true estimate of gas in place. Kaye [48] concluded that calculating gas
in place depending on data about acreage, reservoir pressure and temperature, sand
thickness, porosity and connate water. Brickner [49] described the Bryans Mill field’s
gas as sour gas. In the case of cycling, sweetening sour gas is needed. Tarner et al.
[50] reported that injection of sour gas of Smackover reservoir was stopped due to
corrosion and deposit effects on facilities.
Jacoby and Berry [51] studied the effects of dry gas injection on a volatile oil
reservoir. After injection, a phase equilibrium between volatile oil and dry gas will
be established and lead to oil vaporization. Initially lighter components such as
propane, butanes, and pentanes will be vaporized and after that heavier components
such as the hexanes, heptanes and octanes will be transferred to the gas phase.
Flue Gas. Flue gases contain hydrocarbons (such as methane and propane),
hydrogen, carbon monoxide, carbon dioxide, nitrogen, steam and oxygen [52, 53].
The first successfully system for flue gas injection was installed and put on produc-
tion in 1959 in an oil field in Louisiana state. From then injection of flue gas became
a usual method.
Access to flue gases needs surface plants which carry out the responsibility of
capturing it. Building surface plants need investments and investigation but final NPV
of the project can be higher. However, it can have corrosion effects which corrosion
will be investigated in Chap. 9. This method is a cheap potential gas storage program.
Also, injecting flue gases reduce air pollution [54]. Barstow [55] reported that surface
plant was built in Texas in 1945 in order to inject flue gas into the reservoir. This
project failed due to corrosion. The compressors and relevant equipment were badly
corroded. By the time that equipment broke down, nearly 20 years of injection had
been carried out.
Nitrogen. The most available source of nitrogen is in the air. Air is made up
of a mixture of gases. The major component of air is nitrogen gas, followed by
oxygen. Nitrogen injection has performed widely around the world [9]. In the case
of miscibility, nitrogen should be removed from the natural gas. Nitrogen injection
may have negative effects in some reservoirs. This method increases the dew-point
pressure of gas condensate which causes the early condensate formation [9].
Economic of capturing nitrogen, nitrogen emission, reservoir heterogeneity, the
effect of nitrogen on phase behavior and stock tank liquid composition are the most
important factors in planning a nitrogen injection project [9]. The Cantarell Complex,
Mexico was under pressure maintenance by nitrogen injection since 1996. Pure
nitrogen injection was started at the rate of 1200 MMSCFD which makes this project
the largest nitrogen injection in the world at the time [56].
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 185

Stewart et al. [57] reported that in the Elk Basin field lies in park county, Wyo.
Injected gas composed of 90% of nitrogen and 10% of carbon dioxide which started
in 1949 resulted in 60% of the recovery. This reservoir did not have a gas cap but after
pressure drop under 2128 psi, it was formed. Nitrogen was injected in the crest of
the Ekofisk field in Norwegian sector at 200 MMscf/D. A 3D, three-phase flow with
a capillary pressure function was used to simulate the performance of this fractured
and multiple layered reservoirs [58].
Carbon Dioxide. Carbon dioxide’s emission is a world challenge. CO2 emission
is resulting from the burning of fossil flues and coals. CO2 emissions will be increased
by 35.6 billion metric tons in 2020 [59]. One way for reducing CO2 emissions is to
capture and storage (CCS) it underground. Pressure maintenance can be a way for
(CCS) [60]. Law [61] studied the Athabasca bitumen reservoir in Northern Alberta,
Canada. Pressure maintenance steam-assisted gravity drainage by CO2 and N2 and
flue gas was modelled. CO2 injection leads to bitumen viscosity reduction due to
CO2 solubility. The best recovery is enhanced by CO2 injection. This mechanism
decreases oil viscosity and interfacial tension and has swelling effect. Zheng and Yang
[62] evaluated a three-dimensional (3D) model in a heavy oil reservoirs which have
been discovered in western Canada (Alberta and Saskatchewan). Three horizontal
and five vertical wells were used to inject CO2 to maintain reservoir pressure after
waterflooding. This process increased oil recovery by 8.9–12.4%.
Pressure maintenance using CO2 is reported as a suitable approach to increase
oil recovery of the heavy oil reservoirs by Zheng et al. They studied this method
experimentally and numerically. They also reported that the horizontal wells have a
better performance in comparison to the conventional 5-spot well configuration [63].
Pourhadi and Hashemi studied injection of different gases for enhanced oil
recovery. Their results showed that water-alternating- CO2 injection could signif-
icantly increase the oil recovery due to improve of reservoirs pressure maintenance
and extend the production plateau [64].

5.5 Active Mechanisms During Pressure Maintenance

There are three basic mechanisms to transport miscible and immiscible fluids in
porous media: convection (or bulk flow), molecular diffusion, and mechanical
dispersion. Mechanical dispersion is neglected in the model. A brief description
of convection and molecular diffusion mechanisms follow.

5.5.1 Convection (Bulk Flow)

Convection is the transport of the component as it is carried along within bulk fluid
movement. The driving force for convection (bulk flow) is the potential gradient.
Darcy’s law is an empirical relationship between fluid flow rate in porous media and
the potential gradient:
186 A. Jamili et al.

K kr p A
qp = ∇∅ (5.1)
μp

∇∅ = ∇ p + γ p ∇ D (5.2)

where
q p is flow rate of phase p (oil, gas, or water).
K is absolute permeability.
A is cross sectional area.
kr p is relative permeability of phase p (oil, gas, or water).
μ p is viscosity of phase p (oil, gas, or water).
γ p is specific weight of phase p (oil, gas, or water).
∇∅ is potential gradient.
∇ D is depth gradient.

5.5.2 Molecular Diffusion

Molecular diffusion is the mechanism of a component transport by random molecular


motion. Molecular diffusion is the tendency to mix due to chemical potential gradient.
Bird et al. [65] showed that concentration gradient instead of chemical potential
gradient could be used as the driving force for ideal or near ideal mixtures.
Diffusion is a slow process caused by random molecular motion. But diffusion
can cause convection or bulk movement [66]. Convective flow can have many causes,
such as pressure gradients and temperature differences. However, even in isothermal
and isobaric systems, diffusion can produce convection. CO2 -n-decane system is
a very good example of diffusion causes convection in an isothermal and isobaric
system as experimental work of Chukwuma [67] showed. Therefore, combined mass
transfer flux of diffusion and convection must be used in modelling diffusion:

n i = ρi vi = ρi (vi − v0 ) + ρi v0 = ji + ρi v0 (5.3)

In Eq. 5.3 n i is the combined mass transfer flux (total diffusion flux), ji is the
diffusion mass transfer, and ρi v0 is convective mass transfer. v0 is some convective
reference velocity and should be chosen so that v0 is zero as frequently as possible.
v0 could be mass average velocity (good for constant density liquids), molar average
velocity (good for ideal gases where the molar concentration is constant), or volume
average velocity (best overall for constant-density liquids and ideal gases) [66].
Usually, convective part of mass transfer is neglected, which means that there is
no bulk movement (or the bulk is stagnant) because of diffusion. Basically, for some
cases (ideal mixtures where there is not huge difference in component properties)
neglecting convective part of mass transfer is not a bad assumption, but for non-ideal
mixtures, neglecting convective part of mass transfer may not be a good assumption.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 187

For example, Chukwuma [67] recognized that in the CO2 diffusion process in n-
decane, there is a density change that causes free convection in the system. The free
convection enhances CO2 diffusion and makes diffusion occurs in 45 min while for
other gases it takes several hours. Density of CO2 and n-decane mixture increases
until 70% mol CO2 and then decreases. Diffusion process is modelled based on the
driving force: concentration gradient or chemical potential gradient. The modelling
process will be discussed next.

5.5.2.1 Concentration Gradient Driving Force

Diffusion for ideal mixtures is driven by concentration gradient (Fick’s law):

J A = −cD AB ∇x A (5.4)

where
D AB is binary diffusion coefficient.
C is the mixture concentration.
∇x A is the concentration gradient of component A.

5.5.2.2 Chemical Potential Gradient Driving Force

Diffusion for non-ideal mixtures is driven by chemical potential gradient and is


derived by irreversible thermodynamics as Bird et al. [68]:

⎢ 1
J A = − cD AB ⎢
⎣ x ∇ ln a + [(ϕ A − ϕ B )∇ p − ρω A ω B (g A − gb )]
 A  A
Concentration Diffusion
c RT  
Pressure−Forced Diffusion

+ k T ∇ ln T ⎦ (5.5)
  
Thermal Diffusion

where
a A is activity of component A.
ϕ A = c A V A and ϕ B = c B V B are volume fractions of species of A and B.
c A and c B are concentrations of A and B.
ω A and ω B are mass fractions of A and B.
ρ is the mixture density.
g A and g B are body force per unit mass acting on species A and B.
k T is thermal diffusion ratio.
188 A. Jamili et al.

By dropping pressure, thermal and forced diffusion:


⎡ ⎤

J A = −cD AB ⎣ x ∇ ln a ⎦ (5.6)
 A  A
Concentration Diffusion

Activity is a function of composition and a A = x A γ A , so:



∂ ln a A ∂ ln x A ∂ ln a A
J A = − cD AB x A ∇x A = −cD AB
∂ ln x A T,P ∂ x A ∂ ln x A T,P

∂ ln γ A
∇x A = − cD AB 1 + ∇x A (5.7)
∂ ln x A T,P

If the mixture is “ideal”, then the activity coefficient is equal to unity and the
equation becomes similar to Fick’s law. If the mixture is “nonideal”, then activity
corrected diffusion coefficient is [68]:

∂ ln a A ∂ ln γ A
D aAB = D AB = D AB 1 + (5.8)
∂ ln x A T,P ∂ ln x A T,P

To replace activity with fugacity, we know:




fA
aA = (5.9)
f A0

where


f A is the fugacity of component A in the mixture.


f A0 is the value of the fugacity at standard state.
Also, we know that


dμ A = RT d ln( f A ) = RT d ln(a A ) (5.10)

Substituting Eq. (5.10) into Eq. (5.6):


⎡ ⎤

J A = −cD AB ⎣ x ∇ ln a ⎦
 A  A
Concentration Diffusion

−cD AB x A
= −cD AB x A ∇ ln f A = ∇μ A (5.11)
RT
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 189
  
∂ ln f A ∂ ln x A
J A = −cD AB x A ∇x A
∂ ln x A ∂xA
  T.P

∂ ln f A
= −cD AB ∇x A (5.12)
∂ ln x A
T.P

By comparing Eqs. (5.4) and (5.12), the activity corrected diffusion coefficient is:
  
∂ln f A
D aAB = D AB (5.13)
∂lnx A
T,P

Equation (5.13) shows corrected diffusion coefficient for non-ideality. The diffu-
sive flux of component c in phase p (oil, gas, or water), J cp , is given by Firoozabadi
(2000) as:
 
Jc. p = −ϕ S p ρ p Dc.Mp ∇xc. p + Dc.T p ∇T + Dc.P p ∇ p (5.14)

where
Dc.Mp .Dc.T p .Dc.P p are concentration, temperature and pressure diffusion coefficients
of component c in phase p (oil, gas, or water).
∇xc. p is concentration gradient of component c in phase p (oil, gas, or water).
∇T is temperature gradient.
∇ p is pressure gradient.
ϕ is porosity.
S p is saturation of phase p (oil, gas, or water).
ρ p is density of phase p (oil, gas, or water).
Considering the dispersion term (or correcting the concentration diffusion
coefficients for the effect of porous media) in previous equation, it becomes:
 
Jc.∗ p = −ϕ S p ρ p K c. p ∇xc. p + Dc.T p ∇T + Dc.P p ∇ p (5.15)

where
K c. p is the combined diffusion and dispersion coefficient of component c in phase
p (oil, gas, or water).

5.5.2.3 Molecular Diffusion Coefficients in Hydrocarbons

Molecular diffusion coefficients are calculated by the method of Da Silva and Belery
[69]. This method is based on the published work of Sigmund [70]. From kinetics
theory, the diffusion coefficients for binary systems are related to pressure, tempera-
ture, and composition through the Hirschfelder et al. equation [65], which gives the
low-pressure limit of the density-diffusivity product (Sigmund [70]):
190 A. Jamili et al.

0.5
2.2648 × 10−5 1 1
ρm0 Di0j = + T 0.5 (5.16)
σi2j Ωi j Mwi Mwj

i j and σi j are collision diameter and collision integral of the Lennard–Jones


potential in Eq. (5.16). They are related to the component critical properties (TC,i ,
PC,i , Vc,i , and ZC,i ) of component i through the following set of equations:

1/3 −6/5
σi = 0.1866Vci Z ci i = 1 : nc (5.17)

 
σi j = 0.5 σi + σ j i. j = 1 : n c (5.18)

18/5
εi = 0.1866Tci Z ci i = 1 : nc (5.19)

 0.5
εi j = εi ε j (5.20)

T
Ti j = (5.21)
εi j
1.06036  
Ωi j = 0.1561
+ 0.193 exp −0.47635Ti j
Ti j
 
+ 1.03587 exp −1.52996Ti j
 
+ 1.76474 exp −3.89411Ti j (5.22)

The density-diffusivity product as given by Eq. (5.16) does not remain valid for
the high pressures encountered in hydrocarbon reservoirs. A polynomial correction
expressed as a function of the reduced molar density has to be used as in the following
equation:

ρm0 Di0j
Di j = (0.99589 + 0.096016ρmr
ρm

−0.22035ρmr
2
+ 0.032874ρmr
3
ρmr < 3 (5.23)

ρm0 Di0j
Di j = (0.18839 exp(3 − ρmr )) ρmr > 3 (5.24)
ρm

where
ρm is the mixture molar density.
ρmr = ρρmcm is reduced density of the mixture.
nc 2/3
z i VC.i
and is critical density of the mixture ρmc = 1nc 5/3 .
1 z i VC.i
zi is the mixture composition.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 191

Finally, effective diffusion coefficients for each component of the mixture are
estimated on the basis of the Wilke formula (Sigmund [70]):

1 − zi
De. j = n c zi . (5.25)
j =1 Di j

j = i

5.5.2.4 Molecular Diffusion Coefficients of CO2 in Water

Average diffusion coefficients for CO2 in water have been measured by Unver and
Himmelblau [71] and Thomas and Adams [72] at atmospheric pressure and for a
range of temperatures. Grogan and Pinczewski [73] mentioned that the reported
measurements are well correlated by the Stokes–Einstein equation as follows:

T
DCO2 −Water = 5.72 × 10−12 (5.26)
μw

The constant in the equation is dimensional and the equation therefore requires
2
that T be in K, μ in centipoises, and DCO2 −Water in ms . Grogan and Pinczewski [73]
reported a value of 3.6 × 10−9 ms for DCO2 −Water at 13.1 MPa and 54 °C which is
2

 
close to DCO2 −Oil 5 × 10−9 ms .
2

Campbell and Orr [74] conducted an experiment where CO2 diffused in water
and mobilized the trapped oil in a dead-end pore. Figure 5.2 shows the experiment.
Oil (red Soltrol) was trapped by water at the beginning of the experiment. CO2 flows
in the main channel and has no direct contact with the oil. The CO2 dissolves in the
water and diffuses toward the oil. When CO2 reaches the oil, it dissolves into the oil
causing oil swelling. The swelled oil volume increases and pushes the water toward
the main channel. At the final stage, oil starts to flow in the main channel.

Fig. 5.2 Mobilization of trapped oil by CO2 in water [74]


192 A. Jamili et al.

5.5.2.5 The Effect of Tortuosity on Molecular Diffusion

Tortuosity is a characteristic of a porous medium and defined as the ratio of the true
length of the flow path of a fluid particle and the straight-line distance between the
starting and finishing point of that particle’s motion. Tortuosity depends on porosity
of the porous medium. If the porosity is high, tortuosity is low and vice versa. Because
of the tortuosity in a porous medium, effective diffusion coefficient is lower than their
values without a porous medium. This effect is shown by the following relation:

D
Deffective = (5.27)
τ

where
D is a diffusion coefficient for a component.
τ is tortuosity of the porous medium.
Deffective is the effective diffusion coefficient corrected for porous medium which
should be used in calculations.
Tortuosity is related to porosity through the formation resistivity factor (F) with
the following form:

τ = Fφ (5.28)

where

F = φ −m (5.29)

where m is cementation factor which depends on the nature of porous media and
usually varies from 1.5 to 2.5. Amyx et al. [75] and Langness et al. [76] presented a
good review of the relationship between tortuosity and porosity. They also gave the
following relation based on experimental results:

τ = (Fφ)2/1.67 (5.30)

Substituting Eq. (5.28) into Eq. (5.29) gives one relation between tortuosity and
porosity as:

τ = φ 1.2−1.2m (5.31)

Equation (5.31) is used to estimate the effective diffusion coefficient for oil and gas
components in the porous medium. The tortuosity is often treated as an adjustable
parameter. The tortuosity is used to modify the molecular diffusion coefficients,
adapting it for use in porous media.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 193

5.6 Worldwide Pressure Maintenance Projects

Table 5.1 shows some pressure maintenance project all around the world. In this
table, the reservoir and fluid properties, injection date and injection rate are reported
for each project. It is understandable that pressure maintenance can be applied in
reservoirs with different characteristics.
Gas injection in some Iranian reservoirs was proposed in the 1950s after observing
the decline in production. The aim of the gas injection is for increasing reservoir pres-
sure and consequently increasing the recovery factor to achieve maximum extraction
of oil reserves. In 1976, after conducting studies, gas injection operations in Haftkel
field (Asmari reservoir) began as the first gas injection project in this country. The
second gas injection project was launched in July 1977 in the Asmari reservoir of
Gachsaran oil field. In these two projects, two other oil reservoirs were selected as
sources of gas supply due to the lack of suitable infrastructure for collecting associ-
ated gases and the lack of development of independent gas reservoirs. This led to a
waste of oil in these reservoirs and a reduction in their reserves. Other gas injection
projects in Iran started after these two projects which reported in Table 5.2.

5.7 Reservoir Material Balance for Pressure Maintenance

Material balance which is a well-known calculation for chemical engineers was intro-
duced by a reservoir material balance equation by Schilthuis. Schilthuis presented
equations to relate pressure drop and reservoir’s energy to oil and gas production
[87].
The overall balance below bubble point depends on pore space volume, volumetric
equilibrium fluid properties at reservoir temperature, and uniform pressure with phase
equilibrium temperature and fluid composition. Build-up pressure test gives a fair
average data for the formation.

5.7.1 Muskat’s Material Balance

Consider a porous media that saturated by oil, gas and immobile water which is
shown in Fig. 5.3. The flow equations in the mentioned system are as follow:
Oil phase:

λo ∂ So
∇ ∇( po + ρo gz) = ∅ (5.32)
Bo ∂t Bo

Gas phase:
Table 5.1 Worldwide Pressure maintenance projects
194

Field Injection Injection Formation Formation Porosity Permeability Reservoir Reservoir Oil Oil API References
starting rate thickness depth (ft) (md) pressure temperature viscosity gravity
time (Mcf/D) (ft) (psi) (F) (cp)
Elk Basin 1949 – 250 4538 0.107 91 1327 127 1.4 29 [77]
Tensleep
Hilbig 1933 – – 2000 0.241 570 1228 120 1.86 38 [78]
Bahrain 1939 – 110 2175 0.25 40 1236 140 2.25 [79]
Elk Hills 1976 – 1800 5800 0.236 88 2461 210 0.42 36 [80]
Coalinga Nose 1969 200,000 386 6750 19.9 421 3504 212 0.65 31.9 [81]
Northwes Avard – – 61 4660 0.0758 127 2594 150 0.95 36 [82]
Buena Vista 27B 1949 36,900 43 3900 0.26 250 1577 145 1.02 28.7 [83]
Buena Vista 555 1959 11,900 193 5300 0.23 1063 2288 184 1.2 29 [83]
Canal 1941 11,600 110 8200 0.15 200 3564 210 0.37 37 [83]
Gachsaran – – 1600 7400 0.09 – – – – – [84]
Bibi-Hakimeh – – 1600 5000 0.075 – – – – 30 [84]
Haftkel 1976 400,000 110 2087 Low Low 1100 123 - 37 [85]
Coles Levee 1942 170,200 309 8800 0.195 115 3974 235 0.45 34 [83]
main vestern
Coles Levee 1943 30,400 94 9100 0.205 108 4014 235 0.45 34 [83]
vestern 35
Coyote 1944 21,300 303 5300 0.214 83 2614 185 1.05 33 [83]
Coyama 1950 119,700 141 4500 0.265 600 1661 146 1.6 32 [83]
Cymeric 1948 3300 112 2400 0.336 1300 1210 135 6.2 18.9 [83]
Newhall-Potrero 1944 32,300 192 7000 0.18 75 3129 170 0.72 35 [83]
(continued)
A. Jamili et al.
Table 5.1 (continued)
Field Injection Injection Formation Formation Porosity Permeability Reservoir Reservoir Oil Oil API References
starting rate thickness depth (ft) (md) pressure temperature viscosity gravity
time (Mcf/D) (ft) (psi) (F) (cp)
Rio Bravo 1946 31,000 134 11,400 0.224 500 5014 250 0.2 39.9 [83]
Greely 1948 16,200 131 11,500 0.2 712 5029 252 0.23 37 [83]
West Guara 1958 9560 100 5120 0.29 5000 2280 – 0.58 42 [86]
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs
195
196 A. Jamili et al.

Table 5.2 Pressure maintenance projects in Iran


Field Starting date Injected volume in one year (millionm3 /day)
A 1985 0.02
B 1990 15.12
C 1992 5.61
D 1995 3.78
E 1999 2.12
F 2001 2.01
G 2003 5.1
H 2010 16.41
I 2013 0.11

Fig. 5.3 Schematic of porous media system


R s λo   ∂ Rs So
∇ ρg λg + ∇ pg + ρg gz = ∅ ρ g Sg + (5.33)
Bo ∂t Bo

Capillary pressure:

pc = p g − po (5.34)

Saturation:

So + So + Swc = 1 (5.35)

Oil and gas mobility:


5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 197

ko
λo = (5.36)
μo
kg
λg = (5.37)
μg

By integrating Eq. (5.32) in the absence of gravitational and capillary pressure


over the entire reservoir volume Vr and use divergence theorem.
 
λo ∂ p ∂ So
Ar d Ar = ∅ Vr d Vr (5.38)
Bo ∂n ∂t Bo


where ∂n denotes differentiation along the normal to the reservoir area Ar of the
volume Vr and directed away from the interior of Ar . By Assumption of average
values over the integral, Eq. (5.37) can be written as below.

λo A r ∂ p ∂ So
= ∅Vr (5.39)
Bo ∂n ∂t Bo

The left side of Eq. (5.39) is known as the Darcy’s law for oil phase, so Eq. (5.39)
may write as below.

∂ So
qo = −∅Vr (5.40)
∂t Bo

Equation (5.40) for gas phase is written as follows:



∂ Rs So
qg = −∅Vr ρ g Sg + (5.41)
∂t Bo

The surface gas oil ratio is written as:

qg Bo Bo
R= + Rs = λ + Rs (5.42)
qo B g Bg

At last, by substitution qo and qg from Eqs. (5.40) and (5.41) into Eq. (5.42)
and changing the time variable of differentiation into pressure variable, Eq. (5.43) is
achieved.
 
d So Sg d B g Bg So d Rs λSo d Bo
= fo − + + (5.43)
dp Bg dp Bo dp Bo dp

where

f o = (1 + λ)−1 (5.44)
198 A. Jamili et al.

Equation (5.43) is recognized as Muskat’s material balance equation or the zero-


dimensional approximation of the depletion problem. This equation is also known as
the semi-steady state depletion equation. Since it represents an integral average of the
true flow equation where saturation and pressure gradients are considered uniform
[88].

5.7.2 Integral Material Balance Equations

Muskat’s material balance equation is a linear relation which can be written as the
following relationship.

{E x pansion} − {W ithdrawals} + {I n f lux} = 0 (5.45)

Consider a saturated oil reservoir with an active aquifer where gas-oil contact
remains constant as the injected gas in the gas cap slowly and freely disseminate into
the oil zone, the only region limited to fluid withdrawals. The initial gas saturation
in the oil saturation and the oil saturation in the gas cap is zero and residual water
through the hydrocarbon zone is constant. All the fluids are incompressible. The
material balance of this system based on Eq. (5.45) may be stated as:

Bg − Bgi
{E x pansion} = N (Bt − Bti ) + N m Bti
Bgi

N (1 + m)Bti Swi Btw − Btwi
+
1 − Swi Btwi
N Bti
+ (1 + m)c f p (5.46)
1 − Swi
 
{W ithdrawals} = N p Bo + N p Bg R p − Rs + W p Bw (5.47)

{I n f lux} = G i Bg + We (5.48)

Formation compressibility and two-phase volume factor are defined as follows:

1 V p
cf = − (5.49)
V pi p
 
Bt = Bo + Rsi − Rs Bg (5.50)

By substituting Eqs. (5.46), (5.47) and (5.48) into Eq. (5.45), and the following
material balance equation is obtained:
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 199
   
N p Bt + R p − Rsi Bg − We + W p Bw
N=    B −B  (5.51)
B −B (1+m)B S (1+m)Bti c f p
Bt − Bti + m Bti gBg gi + 1−Swti wi
tw twi
Bt
+ 1−Sw
i i wi i

Rp is the net cumulative gas-oil ratio and Wp is the net cumulative produced water
which respectively includes the injected gas and water [89].

5.7.3 Differential Material Balance Equation

The same unified approach used for integral equation is followed. Differential MBE
is not applicable for under saturated reservoirs and the expansion of connate water
and pore volume will be neglected. The treatment of water influx and gas injection
will be incorporated after the basic MBE is derived.

System E x pansion
   
So d Bo So Bg d Rs m 1 − Swi d Bg Sg d B g
= Vp − + + − (5.52)
Bo dp Bo dp Bg dp Bg dp

Oil Pr oduction
   
So d Bo So Bg d Rs m 1 − Swi d Bg Sg d B g
= Vp − + + − × fo (5.53)
Bo dp Bo dp Bg dp Bg dp

where in the absence of gas reinjection, gravitational and capillary effects:

f o = (1 + λ)−1 (5.54)

If a fraction r of the produced gas is reinjected, the above mobility ratio term must
be replaced as follows:

Bg
λ = (1 − r )λ − r Rs (5.55)
Bo

The rate of oil desaturation is calculated as follows:


   
d So So d Bo So Bg d Rs m 1 − Swi d Bg Sg d B g
Vp = Vp − + + −
dp Bo dp Bo dp Bg dp Bg dp
So d Bo
× fo + V p (5.56)
Bo dp

where the left side of Eq. (5.56) is the rate of oil desaturation. The right side of
this equation is the summation of oil production and oil shrinkage. The final form of
200 A. Jamili et al.

the differential material balance is expressed as follows:



d So Bg d Rs 1 d Bg λ d Bo   1 d Bg
= So + + − (1 + m) 1 − Swi
dp Bo dp Bg dp Bo dp Bg dp

We d B g dWe
+ − (1 + λ)−1 (5.57)
Bg dp dp

where dWdp
e
is water encroachment term directly into the system expansion and We is
the net water influx per unit original hydrocarbon system pore volume. The injected
gas in must be of the dispersed type [90].

5.8 Pressure Maintenance in Fractured Reservoirs

5.8.1 Block-To-Block Process

The block-to-block interaction is governed by fracture permeability, fracture capil-


lary pressure, fracture relative permeability, re-infiltration phenomena, and capillary
continuity. These concepts will be discussed in detail next.

5.8.1.1 Fracture Permeability

There are two fracture permeability types:


Intrinsic Fracture Permeability, Kff . The intrinsic fracture permeability is associ-
ated to the conductivity measured during the flow through a single fracture or through
a fracture network, independent of the surrounding rock (matrix). It is, in fact, the
conductivity of a single channel (fracture) or a group of channels (fracture network).
In this case the flow cross section is represented only fractures void areas (extending
the surrounding matrix area) [91].
Conventional Fracture Permeability, Kf . The intrinsic fracture permeability disre-
gards the rock bulk volume associated to the single fracture; on the contrary, in the
conventional fracture permeability the fracture and the associated rock bulk form a
hydrodynamic unit [91].

5.8.1.2 Fracture Capillary Pressure

Three models are used for fracture capillary pressure as follows:


1: Pcf = 0,
2: Pcf = constant, and
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 201

3: Pcf = Pcf (S w )
The first model has been widely used in numerical simulation, but it seems that
this model does not have physical sound. One may consider zero fracture capillary
pressure when only one phase flowing in the fracture or the pressures of two phases
is equal (like capillary end effect) which can be for very large fracture that may not
existed [92, 93]. The second model is based on the relationship between two flat
parallel plates. The equation for capillary pressure is:

2σ cos θ
Pc f = (5.58)
r

where r is the fracture aperture. The third model assumes that the fracture capillary
pressure curve has a shape similar to that of porous medium but with major differences
in curvature. It is reasonable to expect the fracture capillary pressure to have a very
low irreducible wetting-phase saturation compared with the matrix. The consequence
of saturation-dependent fracture capillary pressure is that full capillary continuity
between stacks of matrix blocks may be expected. Dindoruk et al. [94] proposed the
following expression for the fracture capillary pressure:
 n f
 So − Sor f 
Pc f = Pc0f − σ f ln 
 Pc f ≥ Pc0f (5.59)
1−S or f

where S orf is the residual liquid saturation in the fracture space, Pcf 0 is the fracture
threshold capillary pressure and σf is the logarithmic slope of the fracture capillary
pressure. Pcf 0 is obtained from the adjustment of the initial production data, while
it is estimated by matching calculated and measured data of the entire test period.
Dindourk et al. [94] matched Pcf 0 and σf for every experiment even for similar
experimental setup. They justified using Pcf 0 and σf different and for similar situation
to the scratch of the fracture surface.
Fracture capillary pressure has a significant effect on drainage performance across
a stack of matrix blocks. The flow across a fracture is very sensitive to fracture
capillary pressure due to capillary pressure driving force for liquid film flow [94,
95]. Firoozabadi et al. [95] reported 30 to 40 Psi (207 to 276 kPa) fracture capillary
pressure by assuming fracture surface covered with cones.

5.8.1.3 Fracture Relative Permeability

The relative permeability of the fractured reservoirs is one of the important problems
in studying these kind of reservoirs. Some of the published experimental results
indicate using two relative permeability curves during the simulation of the fractured
reservoirs with wide fractures (Zero capillary pressure). One for matrix and the other
one (two straight lines) for fracture system. In this case, the capillary pressure in
202 A. Jamili et al.

the fracture is zero so, the relative permeability becomes two diagonal straight lines
for the fracture or fracture relative permeability is a linear function of saturation.
For the matrix, the usual relative permeability curves are being used depend on the
wettability of the matrix.
On the other hand, some other publications are based on using the same relative
permeability curve for reservoirs with narrow fractures, which explains the behavior
of the fluid in matrix and fracture. In this case, the capillary pressure of the fracture is
not zero, but the permeability of the fracture is huge. The line with a little curvature
or no curvature will be used for the non-wetting fluid’ relative permeability curve
and a curved line will be used for the wetting phase because, the wetting phase is
continuous and fills the small pores and the other fluid flows through the larger pores.
Edgar et al. [96] believed that generalized Darcy model might not be appropriate
for flow in filled fractures, because the porous medium does not allow the possibility
of ‘blobs’ of one phase transported by another continuous phase. It states that one
phase can only move upon establishing a continuous flow path. Substituting the
equation for permeability of fractures proposed by Witherspoon et al. [97]:

k f = 84 × 106 × w 2f (5.60)

wf is the fracture width.


Into Darcy’s law results in the following equation for water relative permeability
in a fracture:
vw
kr w f = −12μw (5.61)
w 2f ddpx

wf is the fracture width.


μw is water viscosity.
vw is water velocity.
dp
dx
is pressure gradient.

Equation (5.60) shows that different relative permeability curves result from
different pressure gradient in the fracture. When capillary pressure in the matrix
is large relative to viscous force in the fracture, oil is expelled into the fracture at a
rate independent of the viscous forces. The velocity of the water, the oil blob growth,
and the roughness of the fracture wall determine the pressure gradient at which
the blobs are pushed downstream; however most numerical reservoir simulators are
written under the assumption of a Darcy flow model for the fractures. Therefore,
specific relative permeability curves must be obtained for a given injection flow rate
[96]. Because matrix-fracture interaction is controlled by injection rate, it is apparent
that the relative permeability relationships for wetting and non-wetting phases are
rate dependent. Wetting and non-wetting fluids flow simultaneously in both fractures
and the matrix. The combination of these two flow processes results in a combined
relative permeability behavior that has not been determined [96].
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 203

McDonald et al. [98] mentioned that pressure change effects fracture relative
permeability which does not consider in the simulators. Jones et al. [99] showed that
it does whenever flow is laminar in the fracture and relative permeability depends on
fracture aperture, and on roughness of the fracture surface.

5.8.1.4 Capillary Continuity and Re-infiltration Phenomena

In the dual-porosity concept, fracture acts as a conduit for fluid flow and matrix acts
as a source or sink. It means that produced oil from the matrix will flow through
the fracture. In 1979, Saidi et al. [100] observed that the drained oil from the upper
matrix is sucked or re-infiltrated (or re-imbibed) into the lower matrix instead of
flowing through the fracture network. They questioned the validity of the dual-
porosity concept, which assumed that matrix is a discontinuous medium and drained
oil is produced through the fracture only. This phenomenon is called block-to-block
interaction. Prediction of naturally fractured reservoir behavior without considering
block-to-block interaction will be optimistic. From that date, a lot of work has been
done and re-infiltration and capillary continuity terms are used frequently in the
naturally fractured reservoirs literature. First, let’s define capillary continuity and
re-infiltration terms.

Definition of Re-infiltration and Capillary Continuity

Re-infiltration can be defined as a fraction of drained oil from the upper matrix
block that re-infiltrates (re-imbibes) into the lower matrix block, but the re-infiltrated
phase could be a continuous or discontinuous phase. If there is not phase continuity
between matrices, then every block produces separately. Therefore, the effective
height is equal to individual matrix block height and the overall recovery is the same
as individual matrix block recovery.
On the other hand, capillary continuity is an especial case of re-infiltration where
the re-infiltrated phase has to be a continuous phase. Figure 5.4 shows the difference
between capillary continuous and discontinuous cases. As one may see from Fig. 5.4,
phase continuity increases the effective block height during gravity drainage with
respect to the discontinuous case, where each block drains independently. Figure 5.5
shows how re-infiltration and capillary continuity occurs. In Fig. 5.5a, re-infiltration
occurs through contact areas and liquid bridges, where capillary continuity exists,
because in this case re-infiltrated phase (oil) is a continuous phase. In Fig. 5.5b, re-
infiltration occurs by liquid film, but capillary continuity (phase continuity) may not
exist necessary especially toward the end of gravity drainage process, where the oil
drainage rate is small, and it is hard for the re-infiltrated phase (oil) to be continuous.
So it is possible that there is complete re-infiltration between blocks without having
capillary continuity as Saidi [85] mentioned for history matching of Haft Kel field
(Iran). Since the drained oil from the upper block prefers to re-infiltrate into the lower
204 A. Jamili et al.

Fig. 5.4 Continuous and discontinuous matrices [101]

Fig. 5.5 Schematic of re-infiltration and capillary continuity [91]

block instead of producing through the fracture, as observed by Saidi et al. [100],
re-infiltration and capillary continuity delay the oil production.
In re-infiltration, the re-infiltrated phase (usually oil) does not have to be a
continuous phase; therefore, re-infiltration does not necessarily change the ultimate
recovery from a stack of matrix blocks. But in capillary continuity, since the re-
infiltrated phase (usually oil) is continuous, so the ultimate recovery will be increased
significantly from a stack of matrix blocks as one may see from Fig. 5.4. Figure 5.6
shows the definition and difference between capillary continuity and re-infiltration.
As one may see, capillary continuity is a special case of re-infiltration. The problem
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 205

Fig. 5.6 Definition of Re-infiltration and capillary continuity

Fig. 5.7 Constriction


coefficient method [92]

is that these two phenomena overlaps especially during the early life time of a natu-
rally fractured reservoir, which makes it hard to distinguish between them and predict
the reservoir behavior correctly. Some methods have been proposed to analyze and
distinguish between re-infiltration and capillary continuity which will be discussed
next. All methods are similar in terms of assigning saturation to the fracture.
Constriction Coefficients Method. Saidi et al. [102] used constriction coef-
ficients to simulate capillary continuity and re-infiltration processes. Constriction
coefficient between two blocks is defined as:
qc
C= (5.62)
qo

where qo is the total flow from the upper block and qc is the flow through the
contact area or liquid bridges between two blocks (Fig. 5.7). Basically, this method
provides capillary continuity and re-infiltration between blocks by assigning fracture
saturation.
Saidi [103] often assumed capillary discontinuity between a stack of matrix
blocks. He argued that if the fracture aperture (critical fracture aperture) is about
206 A. Jamili et al.

0.002 in. (0.05 mm) or more, capillary continuity between a stack of blocks cannot
be realized. It seems that he found the above number by balancing gravity and capil-
lary forces of a liquid droplet between two flat parallel plates just before detachment.
His model has nothing to do with direct contact areas between two matrixes. The
results of experimental work of Sajadian et al. [104] showed that there is a critical
fracture aperture for stabilizing liquid bridge. They derived the following equation
for critical fracture aperture (tcf ) by equating the weight of the droplet and capillary
force just before detachment as:

1/2  1/2
8σ 8 × 30 × 10−3 mN × 1m
tc f = = 100 cm
1 kg
3ρg 3 × (0.724 − 0.001186) cmg 3 × 9.0806 kg
N
× 1000 g
= 0.336 cm (5.63)

They assumed capillary continuity for fracture aperture less than Eq. (5.63) and
discontinuity for fracture aperture more than Eq. (5.63). Capillary continuity cannot
be described adequately by using Eq. (5.63) for Firoozabadi et al. [105] experiments.
Pcf –krf method. This method treats a fracture as a porous media like the matrix
and assign Pcf and krf to the fracture. Kazemi et al. [106] used fracture capillary
pressure as a matching parameter in reservoir history matching. Saidi et al. [107]
interpreted using fracture capillary pressure as a matching parameter as capillary
continuity between blocks and he wrote: “application of fracture capillary pressure
for better reservoir history match by Kazemi may mean an effective taller block
height”. Basically, fracture saturation will be assigned by using this method; there-
fore, capillary continuity will be achieved between two blocks that surround the
horizontal fracture. Firoozabadi et al. [95, 105], Horie et al. [101], and Dindourk
et al. [94] used this method to investigate the capillary continuity phenomenon. Zero
fracture capillary pressure means there is no capillary continuity between blocks and
each block drains independently [92, 93].
Labastie [108] investigated capillary continuity by changing the contact area
between blocks. He concluded that fracture permeability is the controlling parameter
of capillary continuity between blocks. Festoy et al. [109] modified single porosity
simulator for fractured reservoirs to consider the effect of contact area between
blocks. They showed that even with a small contact area between blocks, the upper
blocks will drain during the life of the reservoir. They assumed that all of the drained
oil from the upper block will be sucked by the lower block. The contact area between
the blocks affects the drainage time schedule. It means with small contact area the
drainage will be longer that if there is larger contact area. However, the ultimate
recovery for small or larger contact areas does not change.
Pseudo-capillary pressure curve. Assigning pseudo-capillary pressure curve
to describe capillary continuity in a stack of matrix blocks is another method. This
method has been used by Thomas et al. [110] and Fung [111] and is based on vertical
equilibrium (VE) concept [112]. Fung [111] included re-infiltration in a computa-
tional grid cell of a dual-porosity reservoir simulation that contains a stack of matrix
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 207

blocks. He first refined the grid into the level of individual matrix blocks, and then
adopted a dual-porosity approach to calculate the drainage performance of the entire
stack (computational grid cell). In this model, the re-infiltration effect is consid-
ered by allowing the communication between each fracture and the matrix block
below. The amount of re-infiltration is characterized by a fractional re-infiltration
parameter. Once the drainage rate versus the average stack saturation is obtained,
the information is then used to construct a pseudo-capillary pressure curve for the
full-scale reservoir simulation. Fung [111] showed a reasonable agreement between
his model and fine grid simulation results. As one may notice, fracture saturation will
be assigned between blocks by constructing a pseudo-capillary pressure curve for the
entire stack of matrix block and this fracture saturation will achieve the continuity
between blocks.
Connection-dependent relative permeability method. Por et al. [113] used
connection-dependent relative permeability between matrix and fracture nodes.
When fracture oil saturation is below threshold saturation, oil will flow to the lower
matrix block instead of flowing through the fracture toward the producing wells. Once
threshold fracture oil saturation is exceeded, then relative permeability will increase
to one linearly. Similar to Fung [111], Por et al. [113] used the dual-permeability
approach to account for capillary continuity. Again, by using this method, assigned
fracture saturation will establish the continuity between every two blocks.
Discussion. Capillary continuity between blocks will be achieved by assigning
saturation in the fracture and maintaining phase continuity between blocks, which
is the common point of the discussed methods. This point will significantly help
understanding and analyzing the re-infiltration and capillary continuity literature.
There are several factors affecting re-infiltration and capillary continuity. Coats
[114] investigated the effect of horizontal pressure gradient in a horizontal fracture
on re-infiltration by using fine grid simulation. Coats [114] calculations showed
no re-infiltration for a fracture with 11° with horizontal plane. Horie et al. [101]
studied the effect of contact points between matrix blocks and showed that there
will be continuity between matrix blocks in the case of direct contact or fine sand
grain between matrixes. Festoy et al. [109] modified single porosity simulator for
fractured reservoirs to consider the effect of contact area between blocks. They
showed that even a small contact area between blocks helps phase continuity (or
capillary continuity), and the upper blocks will drain during the life of the reservoir.
Fung and Collins [111] suggested that when a reservoir shows a high degree
of capillary continuity, the dual-permeability model is more appropriate. In fact, the
effects of partial matrix continuity may be approximately accounted for if an effective
matrix block height that is taller than the actual matrix block height is used. However,
using taller matrix block alters the amount of recoverable oil resulting from gravity.
It seems that pressure drop between matrices is neglected in this model.
Vidal [115] did experiments and showed that the falling film is an important
mechanism in oil recovery from a stack of matrix blocks by gravity drainage. Vidal
[115] concluded that:
208 A. Jamili et al.

– A good knowledge of the history of cycles of saturation and desaturation in the


stack of matrix blocks.
– The accurate description of the block shape and boundary conditions, which affect
the ratio of influence of viscous and capillary forces.
– Reliable and accurate curves of relative permeabilities and capillary pressures for
the successive cycles.
Saidi et al. [100] observed that drained oil from the upper matrix preferred to
travel through the lower block instead of passing along the fractures and by this
way, re-infiltration only delays the oil production by gravity drainage as shown in
Fig. 6 of their paper. Re-infiltration did not change the ultimate recovery from a
stack of matrix blocks. In other words, they did not consider the effect of capillary
continuity which increases the ultimate recovery from a stack of matrix blocks. They
just modified the single block concept. In single block concept, the drained oil from
a matrix is produced, so each matrix drains independently.
They defined the degree of block-block interaction (α) as the fraction of drained
oil from the upper block which re-infiltrate into the lower block and can be varied
between 0 and 1. Figure 2 of their paper is a special case of continuity Eq. (5.46)
that was derived by Cardwell and Parsons [116] for constant instant of time.

∂ Vy ∂ So
= −ϕ (5.64)
∂Z ∂t
From Fig. 6 of Saidi et al. [100] paper, one may expect that if a matrix is supplied by
a certain infiltration rate, the saturation profile does not change until the re-infiltration
rate changes. They investigated the effect of re-pressurizing on oil recovery and
concluded that re-pressurizing could increase oil recovery; however, it will take a
longer time to respond to re-pressurizing in the case of strong degree of block to
block interaction. At the end, they concluded that the actual degree of block to block
interaction could not be estimated from primary performance (because at early stage
the performance is dominated by relative permeability curve and not by block-block
interaction). Good geological and geometrical properties of both the matrix blocks
and the fracture are required.
Firoozabadi et al. [105] studied the re-infiltration process in fractured reservoirs.
They concluded that at early time of production, falling film is the dominant oil
production mechanism, after a while re-infiltration will be dominant mechanism.
However there still be falling film flow which cannot be neglected. However, some
of the oil in the falling film may be re-infiltrating (or re-imbibing) into the matrix.
This phenomenon has not received attention in the literature.
Firoozabadi et al. [117] tried to develop a theory for re-infiltration phenomenon
and add it to the current simulators. They assumed incompressible fluids, infinite
gas mobility and zero fracture capillary pressure. Firoozabadi et al. [117] performed
analytical investigations of the re-infiltration process in vertical and tilted matrix
blocks. They concluded that in a two-phase gas–liquid system, the fracture network
is basically a conduit for the gas and the liquid path is through the matrix blocks.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 209

Tan et al. [118] approximated the drainage rate for all the matrix blocks in a stack
by using only two drainage curves: (1) the drainage curve for the first block where re-
infiltration is absent, and (2) the drainage curve for the rest of the blocks (represented
by the drainage curve of the second block) where re-infiltration can take place. They
assumed that oil drained from the upper matrix blocks will completely re-infiltrate
into the lower blocks, provided the stack is surrounded by gas. After finding the
drainage curve for first and second block by fine grid simulation, the drainage curve
of the stack is obtained by material balance. Also, they assumed infinite gas mobility,
zero fracture capillary pressure and kr o f changes linearly with So f .
Tan et al. [118] scaled drainage curve by using initial drainage rate (maximum rate)
and the average gas saturation of the block approaches Sge at an infinite time to obtain
a similar form of solution for the block saturation history. Actually, they followed
Coats [114] proposal for scaling. They postulated that, under the condition that there
is no fracture capillary pressure, if the matrix blocks are tall or the threshold pressure
is small, a single relationship between qi , S g + and t + can be obtained to approximate
the drainage rate when matrix capillary pressure (due to reservoir pressure) and block
heights vary. They defined qi , S g + and t + as follows:

Sg
Sg+ = (5.65)
Sge
 0 ∗

+ ρgh − Pcm
t =t   (5.66)
ρgh − Pcm 0

 0 ∗

ki ρgh − Pcm hi  
qi = ∗  i qi∗ Sg+ (5.67)
k ρgh − Pcm i h ∗
0

Superscript * refers to the data of the reference matrix block. They concluded that
there are three distinct drainage characteristics curve for each case. Those are:
(1) the first curve belongs to the drainage of the top block where the matrix block
is not subject to re-infiltration, but it is in capillary contact with the second
block,
(2) the second series of curves represent the drainage performance of blocks 2 to
N–1 where both re-infiltration and capillary continuity process take place, and
(3) the third curve belongs to the drainage of the bottom matrix block where
re-infiltration exists, but fracture capillary pressure is zero at the bottom face.
To account for capillary continuity between blocks they used Pcf 0 = 0.0088 psi
and σ f = 0.0023 psi. They showed a good agreement between their results and fine
grid simulation results. It seems that by increasing the number of matrix blocks in
the stack their model will have significant deviation from fine grid simulation.
Horie et al. [119] showed that there would be continuity between matrix blocks
in the case of direct contact or fine sand grain between matrices. Firoozabadi et al.
[105] investigation showed that at early production time there would be continuity
210 A. Jamili et al.

through the matrix blocks and after a while building liquid bridge and detachment will
continue between blocks. However, the rate of building liquid bridge and detachment
will decrease by increasing production time. They concluded that fracture gas/oil
capillary pressure and fracture-liquid relative permeability (the combined effect is
fracture-liquid transmissibility) is the controlling factors of production mechanism.
Saidi et al. [107] argued that the presence and degree of capillary continuity in a
fractured reservoir could be estimated by analyzing by the variation of oil production
rate versus variation of oil column thickness or the variation of oil–water contact.
When oil column thickness reaches a constant thickness for some years, oil produc-
tion either drops sharply or reaches a small rate. It then means either near fully or
negligible capillary continuity. The experimental work of Firoozabadi et al. [105]
verified their statement.
Based on Fig. 5.4, for continuous case, the saturation is continuous function of
block height or it differentiable. For noncontinuous case, the function of saturation
versus block height is not differentiable at fractures, because there is a jump in
saturation. The degree of continuity depends on how much the saturation at the
bottom face of the upper block is close to the saturation at the top face of the lower
block. It is questionable that if a new differentiate function version could be defined
which will be able to determine the degree of continuity.

5.8.2 Gravity Drainage

It is believed that gravity drainage is the most important mechanism for oil recovery
in naturally fractured reservoirs [91, 103], especially in gas invaded zone, and water
invaded zone. Gravity drainage is caused by fluid density difference in the fracture
and adjacent matrix. Gravity drainage mechanism can be categorized as: free-fall
or forced gravity drainage versus equilibrium or non-equilibrium gravity drainage.
In free-fall gravity drainage, gravity difference is the driving mechanism, while in
forced gravity drainage, gravity difference and viscous pressure difference are the
driving forces.
In equilibrium gravity drainage, there is no mass transfer (diffusion) between
oil and gas, but there is mass transfer (diffusion) between oil and gas in non-
equilibrium gravity drainage. Mass transfer occurs by convective transport, molec-
ular transport (molecular diffusion, thermal diffusion, and pressure diffusion), and
mechanical dispersion mechanisms. The combination of molecular diffusion and
mechanical dispersion mechanisms is called hydrodynamics dispersion mechanism,
or convective mixing mechanism.
Mass transfer between oil and gas can affect interfacial tension, capillary pressure
and relative permeability curves and reduce the holdup zone height. Because of these
reasons, gas injection can be very efficient process in naturally fractured reservoirs
(oil swelling). Depending on what type of gas is injecting interfacial tension gradient
can be favorable or not. The differences of capillary pressure lead to flow either from
or towards the fracture which is called “Capillary pumping”. In case of nitrogen
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 211

injection, capillary pumping helps the production from the matrix by decreasing the
interfacial tension in the matrix, while CO2 injection tends to decrease the production
by increasing the interfacial tension in the matrix [120]. Based on simulation, Saidi
concludes that early gas injection can increase oil recovery. Qasem et al. [121] and
Aniefiok et al. [122] simulated gas injection and concluded that early gas injection
is an efficient method, but low matrix permeability oil recovery deteriorates. No
experimental data is available to confirm the increment of oil recovery by early gas
injection.
Miscibility of gas into oil can be achieved by increasing reservoir pressure to
minimum miscibility pressure and in this sense miscible gas injection in naturally
fractured reservoirs is not economic, because a huge amount of gas has to be injected
to increase reservoir pressure to minimum miscibility pressure. Therefore, enhancing
miscibility of gas into oil by economic methods, such as, diffusion and convection
mechanisms is a very important method. Effect of injected gas composition can be
investigated by using single block model [123] to simulate the drainage behavior of
a single matrix block surrounded by fractures by injection of different gases, such as,
N2 , CO2 , dry gas (mainly C1 ), and wet gas into the fracture at the top and producing
oil from the matrix bottom. Also, matrix block height is very important in gravity
drainage behavior. Parachor method can be used to assign interfacial tension to grid
nodes and investigate interfacial tension gradient between the matrix and surrounded
fractures caused by injecting different gases.
Dykstra [124] investigated oil recovery by gravity drainage. He mentioned there
is a certain height in the matrix that above that height gas saturation is below critical
saturation and only oil drains and when gas-oil front passes that height, gas becomes
mobile and there will be two phase flow. Therefore, one may conclude that at early
time production, since only oil drains, fracture capillary pressure could be zero and
when gas-oil front passes that certain height there will be two phase flow and hence
fracture capillary pressure will not be zero. Then oil bridges forming and breaking
takes place [105] and at this stage fracture capillary pressure will oscillate between
zero (breaking bridge moment) and non-zero (forming bridge moment). At the end
only, gas fills the fracture and fracture capillary pressure will become zero again. Of
course, fracture aperture is a key parameter that should be considered. Schechter et al.
[125] reviewed the gravity drainage modelling literature and concludes that there are
four main models for equilibrium gravity drainage process modelling. They are:
1. Cardwell-Parsons-Dykstra (CPD) model (this model will be reviewed in details)
[116, 124]
2. Nenniger-Storrow (NS) model [126]
3. Pavone-Bruzzi-Verre (PBV) model [127] and
4. Luan model [128]
Also, there are empirical models proposed by Aronofsky et al. [129] as:

η = 1 − e−βt (5.68)
212 A. Jamili et al.

Cardwell and Parsons [116] were the first who describe gravity drainage process
with theoretical sound basis. They divided the matrix into two parts. The upper part
is unsaturated region where So is less than 100% and the lower region is 100% oil
saturated. Oil velocity for upper region is:

k dp ∂ So
Vu = ρg − (5.69)
μ d So ∂z

and for lower region (saturated):



kρg H
Vu = 1− (5.70)
μ h0

where H is the height of the top of the 100% saturated region after equilibrium has
been reached and h is the height of the top of the 100% saturated region before
equilibrium has been reached. They ignored capillary pressure and gas pressure to
develop a simple form of continuity equation to solve. They calculated front velocity
as:
   
ρg k1 1 − L−z − kd
H
dz
= (5.71)
dt μϕ (1 − So )

where k1 is the permeability of the medium to the fluid at 100% saturation, kd is the
permeability at the saturation just above the front, L is the draining column length
and z is the front location. Above the front, the following relationship was derived
between z and So :

ρg dk
z= t (5.72)
μϕ d So

To find the initial condition, they differentiated Eq. (5.71) and put it equal to zero
to find the maximum front velocity which will be initial velocity. They assumed that
at zero oil saturation, oil permeability is zero. Dykstra [124] modified this assumption
by defining new saturation variable and assuming zero oil permeability at residual
oil saturation as:
So − Sor
So = (5.73)
1 − Sor

He argued that this definition changed B in the permeability equation:

k = SoB (5.74)
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 213

He found B equal to 3.2, while Cardwell and Parsons [116] found 3.5. Then he
followed their approach and derived front velocity equation as:
⎛$  B %⎞
  μϕ  z  B−1
kr (Si ) 1 − L−z − t
H
dz ρgk ⎜⎜


=  ⎝   B ⎠ (5.75)
dt μϕ μϕ  z B−1
Si − t

Also, he calculated oil recovery as follows:


(L−z)Soi
)z 
Boi
+ So
0 Bo dz z 1 Boi z
So
R =1− L Soi
= − dz (5.76)
Boi
L L Soi Bo 0 Bo

The relation between So and z above the front is:


B−1
1
zμϕ 
So = (5.77)
Bkρgt

Dykstra mentioned that the above equations are valid when S g < S gc . If S g > S gc ,
then gas becomes mobile and the possibility exists that two gas-oil fronts will occur
as described by Martin [130] -one moving up and one moving down. Based on the
author’s knowledge, there is not research work on this idea. Most of research has
been done with dead oil, so it seems that this idea worth looking.
Hasanzadeh et al. studied the gravity drainage in a fractured physical model.
They investigated two types of gravity drainage including free fall gravity drainage
(FFGD) and forced gravity drainage (FGD). They also used nitrogen, carbon dioxide
and air as non-wetting phase and condensate and water as the wetting phase in their
experiments. A detailed sensitivity analysis was performed by them that investigated
the effect of fracture blockage, injection rate, dip angle, and type of wetting phase and
injection fluid on the recovery factor (RF). A numerical simulation was conducted to
simulate the behavior of their experiments. Their results showed that FGD using gas
injection has a better performance than the FFGD. Also, they showed that in the cases
with the variation of the vertical orientation, the RF is decreased. They stated that in
the immiscible gas assisted gravity drainage obtained higher condensate recovery in
comparison to water recovery due to lower interfacial tension and lower viscosity of
condensate [131].

5.8.3 Capillary Hold-up

Gas injection in fractured reservoirs increases oil recovery by gravity drainage, diffu-
sion, and oil swelling. Gravity drainage is a mechanism that would help to recover oil
due to density differences between oil in the matrix and gas in the fracture. However,
214 A. Jamili et al.

Fig. 5.8 Capillary holdup in fractured reservoirs

the performance of gravity drainage is limited by threshold height, which depends


on matrix capillary pressure, and matrix block size. Figure 5.8 shows the concept of
capillary holdup.
A typical capillary pressure versus saturation is shown in left side of Fig. 5.8.
Threshold capillary height (hc ) can be calculated from the equilibrium between
gravity and capillarity as:

Pc = Pg − Po
hc =   (5.78)
ρo − ρg gh

where
Pc is the capillary pressure.
Pg is the gas pressure.
Po is the oil pressure.
ρo is the oil density.
ρg is the gas density.
In order to assure displacement of oil by gas in fracture, block height must be
higher than threshold height (hc ). In the case of low permeable and small matrix block
size, gravity drainage may not be an efficient mechanism. In such cases, diffusion for
small matrix block size such as Ekofisk field (North Sea) or low permeable Asmari
limestone reservoirs (Iran) can be an important recovery mechanism during high
pressure gas injection. Diffusion reduces high interfacial tension which results in
threshold height reduction. However, in case of lean gas injection, IFT will increase
during the process which is the opposite of the first case. As oil in the matrix dissolves
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 215

the injected gas from the fracture, the oil swells and produces into the fracture system.
In fact, in several predominantly oil wet Iranian reservoirs such as “Ahwaz” reservoir
where the height of the matrix blocks is normally between 3 and 5 m, the residual
oil saturation in the water-invaded zone after 40 years remains very high and gas-oil
gravity drainage seems to be an efficient mechanism for releasing the oil from such
reservoirs. The matrix block height in Iranian reservoirs varies between 1 and 20 ft.
Therefore, in some Iranian reservoirs such as “Gachsaran” in which the block height
is less than 2 m, diffusion could be the dominant recovery mechanism.

5.8.4 Diffusion

Few laboratory publications have been devoted to describing the diffusion mechanism
in naturally fractured reservoirs. No laboratory work was published on the diffusion
effects in naturally fractured reservoirs before 1990. During that period (before 1990),
simulation studies were conducted to investigate the diffusion mechanism on oil
recovery in naturally fractured reservoirs. The laboratory and simulation studies of
diffusion as a recovery mechanism in naturally fractured reservoirs are discussed in
what follows. Van Golf-Racht [91] and Saidi [103] were among the firsts that provided
good explanation of molecular diffusion role in naturally fractured reservoirs.

5.8.4.1 Laboratory Studies

Morel et al. [132] conducted laboratory studies of the effect of diffusion in 1-


dimension on oil recovery in naturally fractured reservoirs. Figure 5.9 shows the
layout of the experiments. The experiments were performed with cores of Paris Basin
Chalk (0.105*0.105*1.1811 ft3 ). The permeability and porosity of the samples were

Fig. 5.9 Diffusion


experiment layout [120]
216 A. Jamili et al.

Table 5.3 Description of 1-dimension diffusion experiments by Morel et al. [132]


Experiment Composition Initial gas Pressure Total Gas Injected
no. of the mixture saturation in (MPa) experiment injection gas
(mole %) core (%) time (days) flow rate
in the
fracture
(cm3 /h)
M3 C1 (44.1%), 0 10.1 24 4 then 8 Methane
C5 (55.9%)
M4 C1 (52.4%), 25 10.1 16 4 then 8 Methane
C5 (47.6%)
M8 C1 (45.8%), 7.2 9.8 15 4 Methane
C5 (54.2%)
M5 C1 (52.4%), 29 10.1 16 4 then 8 Nitrogen
C5 (47.6%)
M6 C1 (44.1%), 0 10.1 73 4 Nitrogen
C5 (55.9%)
M7 C1 (50.7%), 0 11.7 13 8, 12 then Nitrogen
C5 (49.3%) 16

2 md and 40%, respectively. Cores were saturated with a binary mixture of C1 –C5 .
Methane or nitrogen was injected in the fracture. They investigated the effects of the
diffusing gas (N2 or C1 ), gas flow rate in the fracture, and initial gas saturation in the
core. The experiments were performed at 38.5 °C. Table 5.3 shows the details of the
experiments. They concluded the following:
1. Initial gas saturation has little effect on oil recovery.
2. Pentane recovery is linear with time, which indicates that the recovery process
in not governed by a pure diffusion mechanism.
3. Pentane recovery by methane injection is 1.6 times faster than recovery by
nitrogen injection at corresponding times. The pentane concentration in the gas
phase in the core was 1.6 times higher for methane injection than for nitrogen
injection.
4. In the nitrogen injection case, saturation profiles along the core revealed a strong
capillary end effect resulting in accumulation of oil in the matrix near the fracture
(Fig. 5.10).
5. When nitrogen is the diffusing gas, the flow rate has a small effect on pentane
recovery; whereas flow rate greatly affects methane production.
Chukwuma [67] studied diffusion of CO2 into n-decane at 100 °F and 206 psia.
Figure 5.11 shows the experimental setup. Glass rods of different diameters (2 mm,
3 mm, 4 mm, etc.) and Pyrex glass beads of 4 mm diameter were used as packing in
the study. CO2 diffuses into n-decane from the top. He recognized that the density of
a CO2 and n-decane mixture had an unusual behavior increasing up to 70% mol CO2
and then decreasing at higher concentrations of CO2 . The density change causes
free convection in the vertical direction with the denser fluid flowing down. Free
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 217

80

Case a (C1-C5)
70
Case b (N2-C5)
60
Gas Saturation (%)

50

40

30

20

10
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Distance from Fracture (m)

Fig. 5.10 Gas saturation profile after 39 days along the core for experiment no. M6 of Table 5.2
[120]

Fig. 5.11 Schematic of diffusion experiment [121, 122]

convection enhances CO2 mass transfer. It took only 45 min to saturate n-decane by
density induced vertical flow while for other gases, such as methane, it takes several
hours. He concluded that the free convection causes the effective diffusivity to be
much higher than a typical molecular diffusion. For example, the asymptotic value
of the effective diffusivity for carbon dioxide in n-decane is about 0.2 cm2 /sec at 206
psia and 1000 F whereas the molecular diffusivity for ethane in n-decane at the same
temperature and pressure is about 5.0 × 10–5 cm2 /sec.
Renner [133] used an experimental setup similar to Chukwuma’s [67] experi-
mental setup (Fig. 5.11) to study CO2 and ethane diffusion into n-decane at 100 °F
temperature and pressures up to 846 psia. As CO2 diffuses into the oil (n-decane),
218 A. Jamili et al.

the pressure tends to drop in the CO2 space. As this occurs, the pressure raised by
compressing the CO2 . From the movement of the piston rod and the linear posi-
tion transducer on the gas metering vessel, the volume of CO2 injected to maintain
constant pressure over the rock face as a function of time may be readily determined.
Because CO2 mass transfer into oil (n-decane) results in swelling of the oil, the
gas/oil interface will move as a function of time. Horizontal and vertical Berea core
(2-in diameter and 6-in long) setups were used in the experiments to investigate the
effect of gravity-induced convection on the observed mass transfer. He observed that
the effective diffusivity of CO2 in n-decane in vertical cores is more than in hori-
zontal cores which appears to be because of combined diffusion and gravity-induced
convection processes. On the other hand, diffusivity of ethane in n-decane is not
affected by the orientation of the core.
Thiebot and Sakthikumar [134] studied gravity drainage and mass transfer in
cylindrical cores surrounded by fractures (Fig. 5.12). They used limestone and chalk
cylindrical core with a length of 40 cm and permeabilities of 60 md and 2 md,
respectively. First, the core was saturated with live oil, representative North Sea light
oil with a bubble point pressure of 180 bar at a reservoir temperature of 132 °C.
Second, equilibrium gas was injected at the top of the core in the fracture and oil was
produced from the bottom. Equilibrium gas is a gas in thermodynamic equilibrium
with the live oil used in the experiment. Therefore, there is no mass transfer between
the equilibrium gas and the live oil. Gravity is the recovery mechanism in this stage.
The step was continued until oil production ceased (gravity drainage equilibrium).
Third, methane or nitrogen was injected instead of the equilibrium gas. Mass transfer
between nitrogen and methane as non-equilibrium gases and live oil in the core occurs
in this stage. They concluded that injection of non-equilibrium gas leads to significant
additional oil recovery even after gravity drainage equilibrium.
Le Romancer et al. [135, 136] performed similar experiments as Morel et al. in
1-D conditions (Fig. 5.9) on chalk cores saturated with a methane-pentane mixture in
the presence of different water saturations and with three diffusing gases: nitrogen,

Fig. 5.12 Experiment setup


of Thiebot and Sakthikumar
[123], Darvish et al. [124],
and Karimaie [125]
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 219

methane, and carbon dioxide. Table 5.4 shows the details of the experiments. Similar
to Morel et al., it was observed that there is an accumulation of oil in the matrix
near the fracture surface when nitrogen is injected. The gas saturation profiles were
similar to Fig. 5.10. Figure 5.13 shows the influence of the diffusing gas type on the
pentane recovery. Based on Fig. 5.13, Le Romancer et al. [135] claimed that only
nitrogen injection allows the obtaining of a constant and high pentane production
rate. Therefore, it was concluded that in a diffusion process it is essential to keep
the highest oil saturation near the fracture. From this point of view, nitrogen is an
interesting candidate. Their carbon dioxide diffusion experiment is simulated in this
dissertation.
Riazi et al. [137] conducted a laboratory experiment to study the diffusion mech-
anism at reservoir conditions (Fig. 5.14). In their experiment, diffusion of N2 into
a mixture of oil and gas (in matrix) at 270 bar and 403 K was studied. The oil
components were N2 , CO2 , C1 , C2 , C3 , iC4 , nC4 , iC5 , nC5 , C6 , C7+ . Cylindrical core
samples (8.3 cm height and 5.1 cm diameter) from the Ekofisk field in the North
Sea were used in a vessel with limited free-volume which was purged with nitrogen
immediately following depressurization from the initial bubble point at 382.8 bar
to 275.9 bar. Porosity and permeability of a core sample were 0.31 and 0.29 md,
respectively. A core sample was supported by the vessel so that all the surfaces were
open to the free-volume. The diffusion process was monitored by analysis of the gas
composition in the free volume with time. Their simulation of the experiment will
be discussed in the next section. The results showed the importance of diffusion in
recovery of oil components.
Le Gallo et al. [120] used the same setup (Fig. 5.9) as Morel et al. [132] to study
diffusion in 1-dimension in Paris Basin Chalk. A description of the experiments is
given in Table 5.5. Le Gallo et al. [120] concluded that capillary phenomenon inside
the matrix contributes to liquid flow towards the fracture and may be enhanced if
interfacial tension is increased by injecting of a gas such as nitrogen.

Table 5.4 Description of 1-dimension diffusion experiment by Le Romancer et al. [135, 136]
Test Injected Water Initial gas Composition Gas flow Pressure Total
gas saturation saturation of the mixture rate in the (MPa) test
(%) (%) (mole %) fracture time
(cm3 /hr) (days)
Mure C1 0 0 C1 = (44%) 4 10.1 23
M6 N2 0 C5 = (56%) 73
M10 N2 30 60
M11 C1 30 39
M12 C1 13 52
M13 N2 13 49
M25 CO2 11 C1 = (28%) 6.3 95
C5 = (72%)
220 A. Jamili et al.

100
CO2/C1-C5
90
C1/C1-C5
Pentane Recovery (initial mass %)

80
N2/C1-C5
70

60

50

40

30

20

10

0
0 500 1000 1500 2000 2500
Diffusion Time (hour)

Fig. 5.13 Effect of diffusing gas type on pentane recovery [126]

Fig. 5.14 Schematic of


high-pressure experimental
cell [137]

Darvish et al. [138] conducted an experiment to study the effect of CO2 injection
into cylindrical cores (60 cm long and 4.6 cm diameter) from North Sea (Maas-
trichtian chalk) surrounded by fractures (Fig. 5.12) at reservoir conditions. Perme-
ability and porosity of the core were 4 md and 44%. The oil components were N2 ,
CO2 , C1 , C2 , C3 , C4 , C5 , C6 , C7+ . The volume between core and core holder (fracture)
was filled by Wood’s metal. After saturating the core with the oil mixture, a fracture
volume surrounding the core was created by heating the solid core and melting the
wood’s metal and draining the melted wood’s metal from the space between the core
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 221

Table 5.5 Description of Le Gallo et al. 1-dimension diffusion experiment [120]


Test Injected Water Initial gas Composition Gas flow Pressure Total test
gas saturation saturation of the mixture rate in the (MPa) time
(%) (%) (mol%) fracture (days)
(cm3 /h)
M5 N2 0 29.5 C1 = (52.4%) 4–8 10.2 16
C5 = (47.6%)
M29 C1 0 0 C5 = (100%) 4 10.2 65
M30 C1 0 0 C1 = (37%) 4 10 95
C5 = (49%)
C15 = (14%)

and the core holder. The oil circulation was continued until fracture and core both
were completely saturated with oil. Once the sealing material from the fracture was
removed, the oil in the fracture was displaced by injecting CO2 at high flow rate.
Then CO2 was injected at the top of the core and oil was produced from the bottom.
The experiment was performed at 300 bars at 130 °C.
The Eclipse compositional simulator was used to simulate the experiment. Mass
transfer between gas in the fracture and oil in the matrix is not considered in Eclipse.
Gas–gas and oil-oil diffusion are allowed in Eclipse only. Therefore, Darvish et al.
[138] had to initialize the fracture with oil and gas phases of rich CO2 to initiate
diffusion between oil in the matrix and oil in the surrounding fractures. The frac-
ture was initialized with a mixture of 95 mol% CO2 and 5 mol% of the heaviest
component. The fluid inside the fracture has a two-phase condition in which liquid
phase has a very high concentration of CO2 . The presence of two-phase condition
in the fracture with a high concentration of CO2 in its liquid phase would start the
liquid–liquid diffusion from the fracture to the matrix. A zero gas and oil diffu-
sion coefficient were assigned for the heaviest component. The simulation results
showed that the key mechanism to recover oil from a tight matrix block is diffusion
and gravity drainage has no significant effect. They recommended that the existing
compositional simulators should be updated to consider gas (in the fracture)-oil (in
the matrix) mass transfer on oil recovery.
Karimaie [139] investigated gas injection (secondary recovery) and gas injection
after water injection (tertiary recovery) in oil-wet carbonate cores. The objective
was to investigate an EOR process for oil-wet carbonate fractured rocks. The core
samples were 20 cm long and 3.8 cm diameter. He used C7 -C1 as oil. Porosity and
permeability values were in the ranges of 8–25% and 1.5 to 130md, respectively.
His experimental setup and procedure were the same as for the Darvish et al. [138]
experiments (Fig. 5.12). Secondary gas injection experiments were done at 220 bars
and 85 °C. In secondary gas injection experiments, equilibrium gas was initially
injected to displace oil by gravity. Equilibrium gas was in equilibrium with the oil in
the core and therefore, there was no mass transfer between the equilibrium gas and
the oil. Once oil production ceased, the second period of pure CO2 or N2 injection
followed. In tertiary gas injection, first oil was displaced by water injection at 220
222 A. Jamili et al.

bars and 85 °C. Then equilibrium gas injection started at 210 bars and 85 °C followed
by the second period of equilibrium gas, N2 or CO2 injection at 220 bars and 85 °C.
He claimed that diffusion plays an important role in both secondary and tertiary oil
recovery. He showed experimentally that tertiary oil recovery increased by increasing
injection pressure from 210 to 220 bar at 85 °C. However, the efficiency of the process
strongly depends on the type of gas. Injecting CO2 resulted in higher recovery than
equilibrium gas or nitrogen injection in tertiary recovery. He claimed this is due to the
fact that, in CO2 injection, several mechanisms such as gravity drainage, diffusion,
swelling, and IFT reduction are contributing to oil recovery. Also, in secondary
recovery, when nitrogen is injected, ultimate recovery is lower than CO2 injection.
No detailed simulation was done.

5.8.4.2 Simulation Studies

Coats [114] included the effect of diffusion in dual-porosity models. Diffusion coeffi-
cients for liquid–liquid diffusion are about 100 times smaller than those for gas–gas
diffusion. Liquid–gas diffusion coefficients are larger than liquid–liquid diffusion
coefficients but still less than gas–gas diffusion coefficients. Therefore, Coats (1989)
neglected gas-oil and oil-oil diffusion between fracture and matrix in his formulation
and only gas–gas diffusion was considered as:
   
Diffusion between matrix and fracture = ϕSgg Dg ρg yi m − ρg yi f (5.79)

where
Sgg is the geometric mean of matrix and fracture gas saturation.
Dg is gas diffusion coefficient.
ρg .ρo are molar densities of gas and oil.
(yi )m .(yi ) f are mole fraction of component i in gas phase in matrix and fracture.
ϕ is matrix porosity.
Da Silva and Belery [69] simulated the effect of diffusion on oil recovery from
highly fractured reservoirs with low matrix permeability in the North Sea and in
Africa. The oil components were C1 , C2 –C6 , and C7+ . The injected gas was nitrogen.
The simulation studies were done at 266 °F and 4415 psia. The maximum matrix
block height was 4 ft in their simulations. The diffusion equation for a matrix block
was solved analytically for a step change in concentration at the matrix boundary.
The analytical solution provided the concentration of each component as a function
of time. Their analytical simulation results showed the significant effect of diffusion
on the oil recovery, especially for small matrix block size of the order of several
feet or less. They suggested considering the effect of diffusion on oil recovery in
simulation of naturally fractured reservoirs.
Thomas et al. [140] conducted a simulation study of nitrogen injection into the
highly fractured Ekofisk field in the North Sea. The model temperature and pressure
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 223

were 268 °F and 4000 psig. Bubble point pressure was 5545 psia. They defined the
diffusion time as the time required to increase concentration of N2 in the core to 99%
by diffusion. They showed that diffusion time for 1 and 10 ft blocks are 10 days and
5 years, respectively. The interfacial tension was increased by nitrogen diffusion.
Hua and Whitson [141] simulated experiment no. M5 shown in Table 5.3. Their
model combined an analytical solution for mass transfer in the fracture with a numer-
ical model in the core. An analytical solution in the fracture was used to define a
mass transfer coefficient between matrix and flowing gas in the fracture. Convection
(driven by pressure gradient) between matrix and fracture is not considered in the
model. They showed that diffusion is an important mechanism for transporting N2
and C1 in the porous media. C5 is transported to the fracture face mainly by oil
convection inside the core. They also recognized the importance of correction of
capillary pressure for the variation of interfacial tension due to gas diffusion in oil
recovery calculations. They used a ternary diagram to explain why pentane recovery
is not only by pure diffusion. They explained that as core fluid contacts nitrogen, the
amount of pentane in the oil phase should increase, which means that pentane will
diffuse from the fracture into the core. This is impossible since the injected gas does
not contain pentane from the fracture. The only way to keep phases in equilibrium
is to have pentane supplied from the lower part of the core. This is the reason behind
oil convection from the matrix towards the fracture.
Fayers et al. [142] simulated experiment no. M5 (nitrogen diffusion experiment)
of the Morel et al. diffusion experiments (Table 5.3) to test their compositional
simulator. The computational mesh had 20 grid blocks along the core and 3 grid
blocks along the fracture, which allowed its inlet, mass transfer region, and outlet
to be represented. The mass transfer coefficient between matrix and fracture was
evaluated using a laminar flow theory similar to that described by Hua and Whitson.
They showed the importance of correcting capillary pressure with interfacial tension
in the calculations. Also, they recognized that the shapes of the calculated saturation
profiles are strongly dependent on the selection of a capillary pressure curve and on
the accuracy of determining variations of interfacial tension.
Riazi et al. [137] solved the diffusion equation (Eq. 5.61) analytically to simulate
their experiments (Fig. 5.14). They treated the fracture as a boundary condition for
the matrix. Two boundary conditions were studied for the fracture-matrix interface.
They were a stagnant condition and high flow in the fracture. Their simulation results
showed good agreement with experimental data (composition of methane versus
time) for both cases. They recognized that diffusion is a very important mechanism
in oil recovery.
Saidi [85] simulated performance of the Haft Kel field at Iran. Matrix block size
varies from 8 to 14 ft in the Haft Kel field. Permeability changes between 0.05 to 0.8
md. He showed the importance of diffusion during history matching of the Haft Kel
reservoir.
Lenormand et al. [143] developed a mass transfer coefficient between matrix
and fracture similar to the Hua and Whitson [141] model. The model was used
successfully to simulate the following experiments:
224 A. Jamili et al.

1. M5 nitrogen diffusion experiment in Table 5.3 [132].


2. M12 methane diffusion experiment in Table 5.4 [135, 136].
3. M29 and M30 methane diffusion experiments in Table 5.5 [120].
Hoteit and Firoozabadi [144] simulated gas injection using finite element methods.
The domain of the model is a 2-D vertical cross-section (xz) with 500 m length
and 100 m height with different fracture spacing of 100 m*10 m, 10 m*10 m, and
10 m*5 m. Matrix permeability was set 1md or 0.1md in the simulation studies.
Matrix porosity was 20%. Fracture relative permeability was linear. Capillary pres-
sures in matrix and fracture were assumed zero. Table 5.6 presents the details of
their simulation study. One injection well and one production well were defined in
the model. The injection well was located on top right corner and the production well
was located at the lower opposite corner. They considered the effect of non-ideality
to calculate the diffusion coefficients in a multi-component mixture. They concluded
that for a low permeability matrix (1 and 0.1 md) the effect of diffusion is much more
than what current models predict. They treated the fracture as a boundary between
adjacent matrices. In their simulation, the pressure, saturation, and mole fraction
in the fracture were calculated by interpolation between adjacent matrices. Their
simulation results showed 25% increase in oil recovery by including diffusion with

Table 5.6 Simulation examples of Hoteit and Firoozabadi [144]


Example no P(bar) at top T (K) Oil Injected gas Gas injection Production
of the model composition composition rate (PV/day)
1 38 366 C3 C1 1.30E−04 BHP =
38 bar
2 320 366 CO2 & C3 , C1 6.80E−05 BHP =
C1 , C2 , C3, 320 bar
C4, C5, C5,
C6, C7–9,
C10+
3 175 366 CO2 , N2 –C1 , CO2 6.20E−05 BHP =
H2 S–C2 –C3 , 175 bar
C4 –C6 ,
C7 –C9 ,
C10 –C14 ,
C15 –C18 ,
C19+
4 437 410 CO2 , N2 –C1 , CO2 –N2 –C1 , 5.76E−05 Constant rate
C2 , C3 , C2 , C3 , = 7.2e–5
C4 –C5 , C4 –C5 , PV/day
C6 –C7 , C6 –C7
C8 –C11 ,
C12 –C19 ,
C20 –C29 ,
C30+
BHP = Bottom hole pressure
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 225

their method relative to the case without diffusion. The effect of diffusion was more
pronounced for smaller fracture spacing.
Alavian et al. [145] simulated a secondary CO2 injection experiment of Karimaie.
In the Karimaie [139] experiments, equilibrium gas was in equilibrium with the oil
in the core and therefore, there was no mass transfer between the equilibrium gas and
the oil. The SENSOR compositional model (single porosity) was used to simulate
the experiment. SENSOR does not have a diffusion mechanism in the single porosity
model. A cylindrical model (single porosity) with 10 grids in radial direction and
51 grids in vertical direction was used to simulate the experiment. The simulation
results showed the following results:
1. Darcy displacement is the dominant recovery mechanism in the Karimaie [139]
secondary experiment during the equilibrium gas injection period because of a
low conductivity in the surrounding fracture. The fracture space in the Karimaie
[139] experiments was created by melting wood’s metal initially filled the space.
Simulation results indicated that the fracture had low conductivity (20–30 md)
which means melting wood’s metal was not a successful process.
2. It was concluded that near-miscible displacement was the dominant production
mechanism during secondary CO2 injection.
3. Gravity-capillary forces had a minor effect in Karimaie [139] experiment.
Moortgat et al. [146] simulated the Darvish et al. [138] CO2 experiment by finite
element methods. Their simulation method is the same as Hoteit and Firoozabadi
[144] method. A Cartesian model with 19*1*40 grids in x, y, and z direction was used
to simulate the experiment. It was found that diffusion was an important recovery
mechanism. However, the impact of diffusion on oil recovery was not as significant
as the Darvish et al. [138] simulation results showed.

5.8.4.3 Multi Component Diffusion

Multi component diffusion occurs when the flux of one component is influenced
by the concentration gradient of a second component [147]. In some cases, the
first component’s flux can be accelerated or decelerated by as much as an order of
magnitude. In other words, the diffusion can cause a temporary unmixing, exactly
opposite to the effect commonly expected. In some cases, the first component can
diffuse against its concentration gradient, from a region of low concentration into a
region of high concentration. Cussler [147] mentioned the lack of experimental data
as the major deficiency to verify multi component mathematical models. Cussler
[147] believed that all available evidence suggests that all mathematical models with
same number of diffusion coefficients give similar results.
226 A. Jamili et al.

5.8.4.4 Diffusion Coefficient

Basically, diffusion coefficients are functions of concentration, temperature, and


pressure. Their functionality of concentration, temperature, and pressure is derived
from non-equilibrium thermodynamics and known as Maxwell–stefan equation. Bird
et al. [68] showed that there are three driving forces for a diffusion process: concen-
tration gradient (molecular diffusion), temperature gradient (temperature diffusion),
and pressure gradient (pressure diffusion).
Firoozabadi et al. [148] studied the role of diffusion in oil composition variation
in fractured reservoirs and they used Bird classification of diffusion driving mecha-
nisms. They claimed that by using their model the oil composition variation can be
explained better, but they didn’t show how. Molecular diffusion coefficients are 2 to
3 orders of magnitude lager than temperature diffusion coefficients and 7 to 9 orders
of magnitude larger than pressure diffusion coefficients.
Firoozabadi et al. [144] used a tensor to represent diffusion coefficients. They
calculated the diffusion flux by multiplying diffusion coefficient by concentration
gradient and because of the effect of multicomponent diffusion some of the off
diagonal diffusion coefficients are negative. Diffusion coefficients in their work are
functions of concentration, temperature, and pressure and unlike their previous work
[148]; they did not follow Bird’s calcification by considering separate concentra-
tion diffusion coefficient, temperature diffusion coefficient, and pressure diffusion
coefficient. Again, the major deficiency of this mathematical model like other math-
ematical models is lack of experimental verification. On the other hand, diffusion
flux of every component is the algebraic summation of that component diffusion flux
in the other components, which for some of them are negative, and will results
to a number and because of this, in the literature, the investigators do not care
about individual component of diffusion coefficient tensor. They care about the total
diffusion flux of that component. Diffusion coefficients are usually obtained from
Sigmund correlations [149], which are made for reservoir conditions. It has shown
that Sigmund correlations [149], are able to predict the diffusion coefficients with
reasonable accuracy.
Da silva et al. [69] extended and used Sigmund correlations to study the effect of
diffusion on the reservoirs at North Sea and Africa. Their simulation results showed
the huge effect of diffusion on the oil recovery and because of that they suggested
studying effect of diffusion carefully and including it in the matrix-fracture fluid
transfer. Darvish et al. [138] experimentally studied CO2 injection into fractured
cores at reservoir conditions. They were able to match their experimental results
using extended Sigmund correlations [149], proposed by Da Silva et al. [69]. Sigmund
correlations have been used in most studies.

5.8.4.5 Field Examples

Table 5.7 shows field examples where diffusion is an important mechanism in oil
recovery [85, 139].
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 227

Table 5.7 Field examples where diffusion is important


Field name Location Matrix block height (ft) K (md) Lithology
Ekofisk North sea <4 2 Chalk Limestone
Haft Kel Iran 8–14 0.05–0.8 Limestone
Gachsaran Iran <6 – Limestone

5.8.5 Dispersion

Dispersion of miscible fluids is caused by several phenomena including molecular


diffusion, velocity profiles in a porous media, stagnant pockets, and variation in flow
paths in a porous media. Diffusion was discussed in the previous section. Mechanical
dispersion represents the effect of porous media structure and it is one of the basic
mechanisms to transport miscible and immiscible fluids in porous media. The effect
of mechanical dispersion is lumped with the diffusion coefficient as follow:
 
*
α L v2p.x + *
αT v2p.y + v2p.z
K c. p.x = M
De.c. +   (5.80)
p.x
ϕ S p v p 

where
Kc,p,x is the dispersion coefficient of component c in x-direction in phase p (oil,
gas, or water).
M
De.c. p.x is the concentration diffusion coefficient of component c in x-direction in
phase p (oil, gas, or water).
+L and α
α +T are the longitudinal and transversal dispersivity factors.
S p is the saturation of phase p (oil, gas, or water).
ϕ is the porosity.
v p.y and v p.z are the velocity of phase p (oil, gas, or water) in y and z directions.
v p is the velocity of phase p (oil, gas, or water).
Dispersion [150] is expressed versus Peclet No.

vd
Pd = (5.81)
Dm

where v, d, and Dm are velocity, mean or effective grain diameter, and molecular
diffusion coefficients.
For Pd < 0.4, the transport process is pure molecular diffusion. For 0.4 < Pd
< 6, the molecular diffusion is of the same order magnitude as that of mechanical
dispersion. For 6 < Pd < 250, the main spreading is caused by mechanical dispersion
combined with molecular. For 250 < Pd < 18,000, the mixing process is dominated by
mechanical dispersion. Discussions of mechanical dispersion can be found in regular
228 A. Jamili et al.

petroleum engineering textbooks such as Enhanced Oil Recovery by Willhite and


Green [151].

5.8.6 Oil Recovery by Gas Injection in Naturally Fractured


Reservoirs: Diffusion or Convection Dominated?

In high permeability rocks or large block dimensions the effect of diffusion is less
pronounced. Criteria will be presented to compare the required time for diffusion
and gravity drainage quantitatively. If the required time for diffusion is much more
than the required time for gravity drainage, then diffusion is unimportant.

5.8.6.1 Diffusion

Coats [114] also solved diffusion equation in 1-dimension (x-direction) to estimate


diffusion transient time. The diffusion equation was derived for a linear horizontal
core with length l initially saturated with fluid of unit concentration and then exposed
to zero fluid concentration at x = 0 and x = l. The diffusion equation in 1-dimension
is defined as follows:

∂ 2C ∂C
= (5.82)
∂x2D ∂tD

Initial condition:

C(x D .t D = 0) = 1 (5.83)

Boundary conditions:

∂C
C(x D = 1.t D ) = 0. (xD = 0.tD ) = 0 (5.84)
∂xD

where
x
xD = (5.85)
(l/2)
Dt
tD = (5.86)
τ (l/2)2

D is diffusion coefficient and τ is tortuosity of the porous medium. This corre-


sponds to the case of a linear core l feet long initially saturated with fluid of unit
concentration and exposed to zero-concentration fluid at x = 0 and x = l. The
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 229

solution for average concentration is derived as:


)l ∞

Cd x , 1
C= 0
=2 e−λn t D (5.87)
l 1
λ 2
n

where
(2n − 1)π
λn = (5.88)
2


Using first term approximation in Eq. (5.87), the time necessary for C to decay
90% from its initial value is:
1 −λ1 t ∗ π
0.1 = e , λ1 = (5.89)
λ1
2 2

The required time is calculated as follows:

0.85τor (l/2)2
t∗ = (5.90)
D
Tortuosity is calculated as [75, 137]:

τor = ϕ1.2−1.2m (5.91)

where m is the cementation factor and usually varies between 0.15 and 2.5. Lagalaye
et al. [152] estimated the tortuosity by using Kozeny-Carman equation (more adapted
to sandstone than carbonates):
-
ϕ
τor = (5.92)
2k(1 − ϕ)2

with k in Darcy.

5.8.6.2 Gravity Drainage

Oil recovery from a single matrix bock is derived by Van Golf-Racht [91] and
will be presented. Gravity and capillarity are the only active forces. Figure 5.15
shows schematic of the matrix block completely surrounded by gas. The boundary
conditions corresponding to this case are:

Z = H, g = g1 (5.93)
230 A. Jamili et al.

Fig. 5.15 A matrix block completely surrounded by gas [79]

Z = 0, o = o2 (5.94)

Interface velocity from oil and gas phases are:

μg U
Φg1 − Φgz = Z (5.95)
K K rg
μo U
Φoz − Φo2 = (H − Z ) (5.96)
K Kr o

By adding these equations and substituting gz − oz = Pc − g(H − Z)ρ:



U μg μo
g1 − o2 − Pc + g Z ρ = Z+ (H − Z ) (5.97)
K K rg Kr o

where

ρ = ρo − ρg (5.98)

o2 = g2 − Pc (5.99)


5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 231

By substituting Eqs. (5.98) and (5.99) into Eq. (5.97) and with Φg1 = Φg2 the
result is:

U μg μo
Pc − Pc + g(H − Z )ρ = Z+ (H − Z ) (5.100)
K K rg Kr o

Or by ignoring the capillary pressure at the exit face (Pc ):

g(H − Z )ρ − Pc
U= μg (5.101)
K K rg
+ (1 − M)Z ]
[M H

where
(μo /K o )
M=  (5.102)
μg /K g

To obtain the relationship between the fraction of the recovered oil and time,
U = ϕ ddtZ is substituted into Eq. (5.82) and when integrated with initial condition t
= 0, Z = 0, the result obtained is:
$ %
Pc Pc
tD.G.Pc = ZD (M − 1) − M + (1 − M) 1 − ln 1 − − ZD
gHρ gHρ
(5.103)

where
Z D = Z /H is the fraction of oil produced, and.
K K rg max gρ
t D.G.Pc = ϕμ gH
t is the dimensionless time.
Equation (5.83) will be used to calculate the gravity drainage time.

5.8.6.3 Sample Calculations

Table 5.8 shows sample calculation results for a 4 ft matrix block with 2 md and 100
md permeabilities. It takes 11.5 years for 2 md matrix block to drain to the threshold
height (25% of the matrix block height), while it takes 0.096 year for the 100 md
matrix block. Diffusion time for the block is 6 years (Table 5.9). As one may see,
diffusion can play an important role when the permeability is 2 md.
Ramirez et al. [153] simulated a matrix block with 20ft*18ft*20ft (x–y-z) dimen-
sions for gas injection. They used 100 md permeability for the matrix block. They
found that the diffusion is unimportant. Table 5.10 shows that it takes 0.2 year
(71 days) for 100md matrix block to drain to the threshold height (60% of the matrix
block height). This time is close to what Ramirez et al. [153] showed in Fig. 9 of
their paper. Diffusion time for the block is 3.91 years (Table 5.11). So, diffusion is
232 A. Jamili et al.

Table 5.8 Gravity drainage


ZD, th 0.25 ZD, th 0.25
time for 2 md and 100 md
permeabilities Porosity 0.3 Porosity 0.1448
H (ft) 4 H (ft) 4
n-Decane 45 n-Decane 45
density (lb/ft3 ) density (lb/ft3 )
CO2 density 1.56 CO2 density 1.56
(lb/ft3) (lb/ft3 )
g (ft/sˆ2) 32.174 g (ft/sˆ2) 32.174
Krg,max 0.95 Krg,max 0.95
K (Darcy) 0.002 K (Darcy) 0.1
CO2 viscosity 0.015 CO2 viscosity 0.015
(cp) (cp)
n-Decane 0.3 n-Decane 0.3
viscosity (cp) viscosity (cp)
M = Oil 200 M = Oil 200
mobility/Gas mobility/Gas
mobility mobility
ZD 0.749999 ZD 0.74999
tDG 850.3869618 tDG 733.5289773
t (Day) 4220.219249 t (Day) 35.14090961
t (Year) 11.56224452 t (Year) 0.096276465

Table 5.9 Diffusion (gas-oil)


Diffusion coefficient, D, cm2 /s 4.00E−05
for a 4 ft matrix block
Tortousity 2.5
H (ft) 4
t (second) 1.97E + 08
t (years) 6.26

unimportant. On the other hand, it takes 5 years for the same matrix block with 2 md
to drain to the threshold height (60% of the matrix block height). As one may see,
diffusion could be the dominant mechanism in this case.

5.9 Well Design and Pattern in Pressure Maintenance

Injection patterns are planned in order to cover and sweep out an expansive area of
reservoirs and make maximum contact with hydrocarbons [9]. Injection patterns are
a function of mobility ratio of fluids, time and cost [154]. The influence of mobility
ratio on the sweep-out pattern in a secondary recovery has been investigated by Dyes
et al. [155], Aronofsky and Ramey [156], Aronofsky [157] and Craig et al. [158].
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 233

Table 5.10 Gravity drainage


P, th (Psi) 2 P, th (Psi) 2
time for 2 md and 100 md
permeabilities (Ramirez et al. ZD , th 0.606771 ZD , th 0.606771272
[153]) Porosity 0.2 Porosity 0.2
H (ft) 20 H (ft) 20
Oil density 35.256 Oil density 35.256
(lb/ft3 ) (lb/ft3 )
Gas density 11.5238 Gas density 11.5238
(lb/ft3 ) (lb/ft3 )
g (ft/sˆ2) 32.15 g (ft/sˆ2) 32.15
Krg,max 1 Krg,max 1
K (Darcy) 0.002 K (Darcy) 0.1
Gas viscosity 0.0238 Gas viscosity 0.0238
(cp) (cp)
Oil viscosity 0.24 Oil viscosity 0.24
(cp) (cp)
M = Oil 10.08403 M = Oil 10
mobility/Gas mobility/Gas
mobility mobility
ZD 0.383229 ZD 0.393128728
tDG 33.46981 tDG 63.0456284
t (Day) 1949.152 t (Day) 71.7028952
t (Year) 5.340144 t (Year) 0.196446288

Table 5.11 Gravity drainage


Diffusion Gas–gas Oil–oil
time for 2 md and 100 md
permeabilities D, cm2 /s 4.00E−04 D, cm2 /s 4.00E−05
Tortousity 2.5 tortousity 2.5
H (ft) 10 H (ft) 10
t (second) 1.23E + 08 t (second) 1.23E + 09
t (years) 3.91 t (years) 39.13

In order to select number and location of injection wells, permeability and porosity
values and heterogeneity, and positions of production wells are needed [159]. By
choosing the wrong injection well pattern, final recovery from the reservoir may be
less than expected [160]. Boatright and Dixon [161] suggested that approximately
10 wells are needed for a 4000 acres field in cycling program. Fay and Parts [162]
modelled invasion pattern of a five-spot cycling project.
In the case of gas-cap injection, all of the injection wells are located around the top
of the structure. The converted production wells may renew their completion because
the perforation needs to replace from the oil zone into the gas cap. The sufficient
number of injection wells depends on the total required injection rate. The rate of each
injection well can be calculated using the inflow performance relationship proposed
234 A. Jamili et al.

by Rawlins and Schellhardt for gas reservoirs [163]:


 2 n
Q = C Piw − Pe2 (5.104)

where Piw and Pe are the bottom-hole injection pressure and the pressure at the
external limit of the gas zone.

5.10 Simulation of Pressure Maintenance

5.10.1 Methods of Fractured Reservoirs Modelling

In fractured reservoir models the reservoir is discretized into two sets of grid blocks
located in the same space, one called the matrix and the other called the fracture.
The matrix continuum is assumed to be comprised of matrix blocks, which are sepa-
rated spatially by fractures. The dimensions of these matrix blocks can be variable
throughout the reservoir and are a function of the fracture spacing, orientation and
width. Within the computational grid, the matrix- fracture transfer is represented by a
single flow term. The flow of fluids through the reservoir occurs through the fracture
system, which is the primary conduit for fluid flow. Wells in a fractured block are
always connected to the fracture.
Frequently, a fine grid simulation [164] (explicit discretization method) is used
to evaluate the accuracy of a proposed transfer calculation. In such a simulation,
the matrix blocks are discretized into several grid blocks. The adjacent fracture may
also be discretized. The idea is that a fine-grid simulation can accurately represent
the interplay of recovery mechanisms on the local scale and does not involve the
assumptions of how matrix-fracture transfer ought to be calculated.
There are several commercial models available to conceptualize and model
naturally fractured reservoirs. These models are classified as dual-porosity, dual-
porosity/dual-permeability, and dual continuum, approaches.

5.10.1.1 Dual Porosity

Fluid flow through the reservoir takes place through the fracture network. The matrix
blocks essentially act as source and sink terms. Figure 5.16 shows the schematic of
the dual-porosity concept.
The fracture medium is a continuous medium, while matrix is a discontinuous
medium. Drained oil from the matrix to the fracture will produce. In this model frac-
ture permeability is much higher than matrix permeability. Dual-porosity model does
not consider re-infiltration and capillary continuity processes. Kazemi et al. [106]
first derived the governing flow equations for black-oil using in a simulator. Odeh
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 235

Fig. 5.16 Dual-porosity model

[165] compared gas injection results of three-dimensional black-oil reservoir simu-


lation problem by different simulators. Behie [166] solved gas injection problems
by black-oil model. Bansal et al. [167] simulated a strongly coupled, fully implicit,
three dimensional, three phases pressure maintenance by gas injection problem.
In standard dual-porosity model [106], all the matrix blocks in a grid block will
be represented by a single matrix node and one matrix-fracture transfer term which
are questionable. The black-oil governing equations are as follows (CMG IMEX
manual, Appendix A):
For the fracture:
Oil:
   
Toxf pon+1
f − γ x
of D + T x
om f p n+1
om J − p n+1
of
n+1 n 
Vb φ So φ So
+ qon+1 − − =0 (5.105)
t Bo f Bo f

Water:
   
Twx f pwn+1
f − γ x
wf D + T x
wm f p n+1
wm J − p n+1
wf
n+1 n 
Vb φ Sw φ Sw
+ qwn+1 − − =0 (5.106)
t Bw f Bw f

And for matrix sub block k:


 n+1   n+1 
x
Tαm k−1/2
pαm k−1 − pαm
n+1
k
+ Tαm
x
k+1/2
pαm k − pαm
n+1
k
236 A. Jamili et al.
n+1 n 
Vb φ Sα φ Sα
− x − =0 (5.107)
t Bα m Bα m

α can be oil or water.


It is apparent that the proper representation of matrix-fracture exchange is an
important aspect of the dual-porosity model [106]. To improve the handling of gravity
and capillary effects on the imbibition and drainage behavior, dual-permeability
model, gravity segregated model, subdomain model and pseudo function model were
proposed.

5.10.1.2 Dual Permeability

The dual-permeability model assumes two superimposed continua: matrix and frac-
ture, which both of them are continuous. Now, fluid flow occurs through both the
fracture network as well as through the matrix blocks. The dual permeability option
can be important for cases where there is capillary continuity and re-infiltration.
Usually, only the vertical k direction matrix–matrix transfer is important. Figure 5.17
represents the dual-permeability concept. In this model since fracture permeability
relative to matrix permeability is not high, both matrix and fracture have equations.
If the fluid potential in the fracture larger than matrix, fluid will flow from the frac-
ture to the matrix, which is not possible in dual-porosity model. The governing flow
equations for matrix and fracture are as follows:
Matrix:
 n+1   
ΔTom x
Δpom − γom x
ΔD + Tom x
f po f − pom
n+1 n+1

Fig. 5.17 Dual-permeability model


5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 237
n+1 n 
Vb ϕ So ϕ So
− − =0 (5.108)
Δt Bo m Bo m

For flow in the fracture the same equation holds with appropriate symbol for
fracture. Fung and Collins [111] suggested that when a reservoir shows a high degree
of capillary continuity, the dual-permeability model is more appropriate. In fact, the
effects of partial matrix continuity may be approximately accounted for if an effective
matrix block height that is taller than the actual matrix block height is used. However,
using taller matrix block alters the amount of recoverable oil resulting from gravity.
It seems that pressure drop between matrixes is neglected in this model.

5.10.1.3 Gravity Segregated Model

This model uses the representative matrix block concept for calculating gravity effects
between the matrix block and the associated fracture. The gravity drainage mech-
anism is represented by a gravity head, γα Dα , in the matrix-fracture transfer
potential [164].

(ϕf − ϕm )α = (pf − pm )α + γα Dα (5.109)

where
  
γog Dg = γo − γg Sg∗ f − Sgma

Lz (5.110)

 
γow Dw = (γw − γo ) Sw∗ f − Swma

Lx (5.111)

γgow Do = −γog Dg − γow Dw (5.112)

Sα∗ f and Sαma



are the normalized saturations. The matrix-fracture transfer function
becomes:

ταmaf = σVb (1 − ϕf )λα (ϕαf − ϕαma + fs γα Dα ) (5.113)

f s accounts for the predominant vertical liquid flow of the gravity drainage process
and is:

f s = k x /L 2x /σ (5.114)

The gravity segregation calculation assumes complete phase segregation in both


matrix and fracture. The interplay of capillary and gravity forces is not treated prop-
erly in this model. Capillary continuity and re-infiltration processes are not accounted
for.
238 A. Jamili et al.

Fig. 5.18 Subdomain model

5.10.1.4 Subdomain Model

This model is a variation of the standard dual porosity option. It allows the user to
refine the matrix blocks in the vertical direction in order to more accurately represent
the gravity drainage process from matrix to fracture, which is essentially vertical.
Each matrix block is refined by the number of subdivisions specified in order to repre-
sent more accurate the fluid pressures and saturations within the matrix blocks. This
resolution is not needed in the fracture, therefore only the matrix blocks are refined.
Figure 5.18 shows this model. This model does not consider capillary continuity and
re-infiltration between matrix blocks.

5.10.1.5 The Pseudo-Function Method

Typically, the matrix-fracture transfer description is generated by fine-grid simu-


lations of a single matrix block or by some analytical considerations of the local
recovery processes under varying or constant fracture boundary conditions. The
transfer calculation is then carried out using pseudo capillary pressure as a func-
tion of the ‘average’ conditions in the matrix and the fracture. Hence, the interplay
of various recovery processes is assumed to have been captured correctly in these
pseudos and the local recovery phenomenon is not solved directly during the simu-
lation. This approach is computationally efficient and can be quite effective if the
pseudos are representative of the local processes from history match through to the
prediction phase.
Kazemi et al. [168] pointed to the potential danger of using pseudo functions based
on the single matrix-block concept and compare their results with fine-grid simulation
for a stack of matrix blocks where re-infiltration is an important phenomenon. Their
fine-grid results show significantly lower drainage rates. Fung [111] used pseudo
functions to incorporating re-infiltration between matrix blocks in dual-porosity
simulators.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 239

Fig. 5.19 MINC model

5.10.1.6 MINC (Multiple Interacting Continua)

Another extension of the standard dual porosity concept is the multiple interacting
continua (MINC) approach. Pruess et al. [169] first presented MINC idea. The main
advantage of MINC is the representation of the matrix-fracture transfer calculation
using a nested discretization of the matrix blocks. This allows a very efficient repre-
sentation of the transient fluid regime, which is often neglected, in the standard dual
porosity model. The nested discretization is one-dimensional and can represent the
pressure, viscous and capillary forces. The gravity force however is not considered
in this nested, one-dimensional matrix refinement. Figures 5.19 shows the schematic
of MINC model.

5.10.1.7 Single Block Approach

The approach used to model naturally fractured reservoirs on a local scale is to


consider a fractured porous media as a single matrix block with an adjacent fracture.
The fracture acts as a boundary condition for the matrix. This approach is a fine-scale
representation of a naturally fractured reservoir since it allows one to study the fluid
flow between the fracture and the matrix block. This approach is known as single
porosity, dual-continuum, and single block model. Figure 5.20 shows the layout of
the model. The matrix is discretized, but the fracture is not, because the fracture
acts as a boundary condition for the matrix. The single-block model has been used
in many research studies. Yamamoto et al. [123] used this model to study a single
matrix block under several boundary conditions.
240 A. Jamili et al.

Fig. 5.20 Single block


approach

5.10.1.8 Hierarchical Fracture Model (HFM)

Simulation of multiphase flow in a fractured porous media is computationally


demanding. Petrophysical properties such as permeability are highly heterogeneous
and change over several order of magnitudes. This computational challenge is moti-
vated for development of fracture models. Hierarchical fracture model (HFM) simu-
late multiphase flow in naturally fractured reservoir with multiple scales and on
stratigraphic and unstructured grids.
The HFM models the fracture explicitly as the main path for fluid flow. Fracture
and matrix properties are defined separately which have the benefit of independent
definition for matrix and fracture grid. Heterogeneous rock (matrix) and highly-
conductive fractures defined on independent grids. After defining matrix and fracture
properties, both medium will be put together in a new geometry [170].
This Model also known as embedded discrete fracture model (EDFM) and extends
the newly developed multiscale restriction smoothed basis (MsRSB) method for
computing multiphase flow in a fracture system. In this method, matrix is represented
in a fine-scale volumetric grid and the fracture in a lower-dimension grid. Then, the
matrix and fracture basis functions and properties are restricted smoothing [171].

5.10.2 Simulation of Naturally Fractured Reservoirs


by Finite Element Method

Finite element method has been used beside finite difference method to simulate
naturally fractured reservoirs. Finite difference method (FD) is simpler and easier to
apply than finite element (FE) method, but FD cannot be applied for irregular matrix
shape. Firoozabadi et al. [101, 172, 173] used FE method to simulate naturally frac-
tured reservoirs. Firoozabadi et al. [101, 172, 173] used mixed finite element (MFE)
and the discontinuous Galerkin (DG) methods to simulate naturally fractured reser-
voirs. Generally, fractures have very small aperture, and therefore volume relative to
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 241

Fig. 5.21 Typical control volumes containing a fracture with the cross-flow equilibrium in the
computational control volume Firoozabadi et al. [90, 161, 162]

matrix size, which places time step instability. Thus, the integral of governing equa-
tions over the fracture can be simplified by dropping one of the spatial variables.
Consequently, fractures can be represented by the edges of the control volumes in
the grid. Based on cross-flow equilibrium concept (similar to vertical equilibrium),
Firoozabadi et al. [101, 172, 173] assumed that the state unknowns (pressure and
composition) in the fracture and in the adjacent cells are the same (see Fig. 5.21).
This assumption is valid as long as matrix elements are not too large. This assumption
removes the necessity of calculating matrix-fracture flux; only matrix–matrix and the
fracture-fracture fluxes are required. They used direct solver package UMFPACK of
Davis and Duff to solve the linear system of equations.
They neglected capillary pressure and diffusion processes. Also, they assumed that
there is no volumetric accumulation in the fracture nodes. Their work is in 2-D and
it seems that for 3-D the simulation will be much more complicated if one may want
to consider the effects of capillary pressure and diffusion processes. They claimed
that mixed finite element (MFE) and the discontinuous Galerkin (DG) methods need
less CPU time and have less dispersion than finite difference method.

5.10.3 Compositional Numerical Simulation

Compositional simulators were developed to predict the phase and compositional


behavior of reservoirs fluids under gas injection. In compositional simulation, it
is assumed that water and hydrocarbon phases are insoluble. Therefore, separate
mass conservation equations are written for the water and hydrocarbon components.
Compositional simulators are written in moles instead of mass, since phase behavior
242 A. Jamili et al.

equations are expressed in moles. For compositional multiphase flow, three forces
must be properly accounted for: viscous, gravity, and capillarity. In addition, if gas
is injected, diffusion mechanism must be included to quantify for mass-transfer
between phases. Details of the model including mass-transfer mechanisms, governing
equations, boundary conditions, initial conditions, and the numerical solution will
be presented in this section.

5.10.3.1 Governing Equations

A compositional reservoir simulator consists of a set of partial differential equa-


tions with appropriate initial and boundary conditions. The equations governing
compositional multiphase flow in porous media arise from three sources [174]:
1. Material balance equations govern transport for each component c in oil and
gas by the convection and diffusion mechanisms. Therefore,
Hydrocarbon components and CO2 ,
 −
→ → 
− −→ → 

kk
∇. ρo xc kkμroo ∇ po − γo ∇ D + ρg yc μrgg ∇ pg − γg ∇ D
    
+∇. φρo So Dc,o ∇xc + φρ g Sg Dc,g ∇ yc
c = 1, 2, . . . , n c

  + q D, f m,c + qC, f m,c
= ∂t φ ρo So xc + ρg Sg yc
(5.115)

The first bracket represents the convection mechanism in oil and gas phases.
Diffusion mechanism is represented in the second bracket. q D. f m and qC. f m are
diffusion and convection mass transfer between matrix and fracture at the matrix-
fracture boundary. One mass balance equation describes water movement by the
convection mechanism only, because the hydrocarbon phases are assumed to be
insoluble with the water phase.
Water phase,

kkr w ∂
∇. ρw (∇ pw − γw ∇ D) = [φρw Sw ] (5.116)
μw ∂t

2. Phase equilibrium between hydrocarbon phases is expressed in the form of


equality between the fugacity (from Peng-Robinson EOS) of each component
in both oil and gas phases,

f o,c = f g,c c = 1, 2, . . . , n c (5.117)

3. Constraint equations that require the phase saturations to sum to unity and mole
fraction in each phase to sum to unity. Besides, it is necessary to relate water,
oil, and gas pressure, that is, capillary pressure relationships.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 243

So + Sg + Sw = 1 c = 1, 2, . . . , n c (5.118)

,
nc ,
nc
xc = 1, yc = 1 c = 1, 2, . . . , n c (5.119)
c=1 c=1

pcog = pg − po , pcow = pg − pw (5.120)


 
ref
If gas-oil capillary pressure pc is reported at a reference interfacial tension
(σr e f ), in some cases it must be corrected by local interfacial tension (σ ) at each grid
as:
σ
pcog = pcog
ref
(5.121)
σr e f

The equations governing compositional multiphase flow in porous media are given
by Eqs. (5.96) to (5.102). This equation system consists of set of (2nc + 6) equations
with the same number of unknowns. The (2nc + 6) unknowns are (po , pg , pw , So , Sg ,
Sw , x1 , x2 , …, xn , y1 , y2 , …, yn ).
Initial and boundary conditions
Initial conditions define the pressure, saturation, and composition distribution at time
equal to zero. Boundary conditions specify the ways in which the reservoir interacts
with its surrounding. The initial and boundary conditions will be presented next.
Initial conditions
It is assumed that there is gravity equilibrium in the model at time equal to zero. Also,
pressure and composition at a reference are known. Since there is gravity equilibrium
at time equal to zero, convective flow vanishes. Therefore, from Darcy’s law:

kkr p  
ρp ∇ pp − γp∇ D = 0 p = gas, oil, and water (5.122)
μp

For a horizontal plane ∇ D = 0, so Eq. (5.103) simplifies to:

∂pp
=0 p = gas, oil, and water (5.123)
∂x
∂pp
=0 p = gas, oil, and water (5.124)
∂y

Equations (5.123) and (5.124) state that pressure, composition, and saturation
are constant in a horizontal plane at time equal to zero. For the vertical direction,
Eq. (5.122) becomes:
244 A. Jamili et al.

∂pp
− γp = 0 p = gas, oil, and water (5.125)
∂z

Equation (5.125) means that vertical pressure distribution is given by the column
weight. Integrating Eq. (5.125) results in:

p p = pr e f + γ p (z − z ref ) p = gas, oil, and water (5.126)

where γ p is the average specific weight of phase p between z and zref height. If
pressure at a reference height is given, then pressure at any point in the model can
be calculated from Eq. (5.126).
Boundary conditions
There are several types of boundary condition. Most of boundary conditions are
discussed in reservoir simulation textbooks. Two types of boundary conditions will
be discussed next: (1) sealed boundary and (2) matrix-fracture boundary.
Matrix sealed boundary conditions
The total mass flux for all components in all phases vanishes at these boundaries.
That is convection flux at the boundary:

kkr p  
ρp ∇ pp − γp∇ D = 0 p = gas, oil, and water (5.127)
μp

Diffusion flux at the boundary:


 
φρ p S p Dc, p ∇xc p = gas and oil c = 1, 2, . . . , n c (5.128)

Equations (5.127) and (5.128) are defined in the model by setting the transmissi-
bilities equal to zero at the sealed boundaries.
Matrix-fracture boundary conditions
The continuity equation in the fracture includes mass transport by diffusion and
convection mechanisms in a laminar flow regime. For example, steady state conti-
nuity equation for a horizontal fracture can be expressed by the following partial
differential equation:

∂ yc ∂ 2 yc
v − De.c 2 = 0 c = 1, 2, . . . , n c (5.129)
∂x ∂z

where v is the average gas velocity over the cross-sectional area normal to the bulk
flow in the fracture, and De.c is the effective diffusion coefficient for component c
between matrix and fracture. The boundary conditions are:
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 245

z = 0, yc = yc,m f
z = H, ∂∂yxc = 0 c = 1, 2, . . . , n c (5.130)
x = 0, yc = yc, f

where
yc,m f is composition of component c in the gas phase at matrix-fracture boundary,
yc, f is composition of component c at the entrance of the fracture, and.
H is the fracture thickness in the z-direction.
Gas stream velocity and physical properties in the fracture are assumed constant
in deriving Eq. (5.129). An analytical solution was derived for Eq. (5.129) to
find composition (yc ) distribution in the fracture. The details of the derivation are
presented by Hua and Whitson [141] and Lenormand et al. [143]. The diffusion mass
transfer rate at matrix-fracture surface was found by differentiating yc as follows:

∂ yc
q D, f m,c = Aρg De,c c = 1, 2, . . . , n c (5.131)
∂z z=0

where ρg is gas stream density in the fracture. After simplifications, the final diffusion
mass exchange rate between matrix and fracture is defined in the model as:
 
q D, f m,c = kc W H yc. f − yc.m f c = 1, 2, . . . , n c (5.132)

De.c  π 2
kc = αs ρg v exp − l −1 c = 1, 2, . . . , n c (5.133)
v 2H

where
W is the fracture width in y-direction,
αs is a factor for considering skin-effect at matrix-fracture boundary, and.
l is the fracture length in x-direction.
If the matrix is saturated with oil (no initial gas saturation presents), then diffusion
mass transfer occurs between flowing gas in the fracture and oil in the matrix. In this
case, yc.m f in Eq. (5.132) is defined as

yc.m f = K c xc.m f (5.134)

where
K c is the equilibrium ratio of component c.
xc.m f is composition of component c in the oil phase at matrix-fracture boundary.
246 A. Jamili et al.

Diffusion between the matrix and fracture is modelled by introducing a source/sink


term (Eq. (5.132)) in the flow equations for the first matrix grid cell adjacent to the
fracture. After a certain amount of each component has entered or left the first
grid during a time step, flash calculations are performed to distribute the entered
or remained amount of each component between oil and gas phases. Mass transfer
between grids in the matrix occurs by diffusion and convection in both oil and gas
phases (Eq. (5.115)). Convection between a matrix grid cell and the adjacent fracture
qC. f m is defined in the model based on Darcy’s law as:
  
q D,m f,c = ρo xc kkμroo (∇ po − γo ∇ D)matri x − (∇ po − γo ∇ D) f ractur e
  .matri x    /
kkrg
+ ρg yc μg ∇ pg − γg ∇ D matri x − ∇ pg − γg ∇ D f ractur e
matri x
c = 1, 2, . . . , n c
(5.135)

If the pressure in the oil and/or gas phase in the matrix grid cell is more than
fracture gas pressure, all components in oil and/or gas will flow from the matrix
to the fracture. On the other hand, if gas pressure in the fracture is more than gas
pressure in the matrix grid cell, then the gas flows from the fracture to the matrix if
difference between gas pressure in the fracture and oil pressure in the matrix grid
cell exceeds threshold capillary pressure. If there is no gas saturation in the matrix
grid (IFT = 0), then the threshold capillary pressure remains constant until a gas
saturation forms in the matrix grid cell (IFT > 0) and the critical gas saturation is
reached. After developing gas saturation in the matrix grid cell, threshold capillary
pressure is scaled with interfacial tension. Oil does not flow from the fracture to the
matrix, because there is no oil phase in the fracture.

5.10.3.2 Numerical Solution

The differential equations governing compositional multiphase flow in porous media


are presented in the previous section. Some of these equations are nonlinear. The
numerical technique replaces all derivatives with the finite-difference approximations
resulting in a set of nonlinear algebraic equations. Then, the resultant equations are
linearized and solved by the iterative Newton–Raphson method.
Basically, four numerical methods have been proposed in the literature to solve the
compositional modelling equations set: (1) Fully Implicit (FI), (2) Implicit Pressure
and Saturation Explicit Composition (IMPSEC), (3) Implicit Pressure Explicit Satu-
ration and Composition (IMPESC), and (4) Sequential Implicit (SI). Fully Implicit
is the general numerical solution technique and other techniques are special cases of
Fully Implicit technique. Young et al. [175] divided solution methods to “Newton–
Raphson” [176, 177] and “Non-Newton–Raphson” [178–181] methods based on the
manner in which a pressure equation is formed. In the Newton–Raphson method
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 247

the iterative technique specifies how the pressure equation is formed. In the Non-
Newton–Raphson method, the composition dependence of certain terms is neglected
to form the pressure equation. These methods will be discussed briefly next.
Fully Implicit (FI). This method is unconditionally stable and has been used in
the literature only for especial cases, because it is too expensive and requires a huge
amount of computer memory for high number of components [176, 182–185]. Coats
[176] reported numerical dispersion using FI method. Thiele et al. [186] showed that
Coats [176] FI model requires substantial storage and computational time than other
methods. Although, this method allows using big time steps which may seems to be
an advantage at first glance by reducing the required time steps, but more iterations
are required for bigger time steps.
Implicit pressure and Saturation and Explicit Composition (IMPSEC). This
method is also called IMPSAT (Implicit Pressure and Saturation) [187, 188], ECIPS
(Explicit Composition Implicit Pressure and Saturation) [189], IMPIS (Implicit Pres-
sure and Saturation) [190], and Semi-Implicit [191, 192] method in the literature.
Quandalle et al. [190] introduced this method. In this method, three equations for
one pressure (Po ) and two saturations (So , Sg ) for each node will be solved implicitly.
For each node, the system to be solved implicitly is 3*3 (one pressure (Po ) and two
saturations (So , Sg )) block matrix in the form [189, 191, 192].
Cao et al. [187] and Haukas et al. [188] derived stability criteria for this method.
Cao et al. [187] showed that this method is more stable and 50% faster than IMPES
method (time step size up to 10 times larger than what is possible with IMPES model)
due to its improved stability and less expensive than fully implicit technique and is
a very suitable technique for problems that aren’t difficult and required high level
of numerical stability. Cao et al. [187] claimed that strength of coupling is different
from grid-block to grid-block, pressure is always strongly coupled then saturation
and composition are less strongly coupled. In IMPSAT, only pressure and satura-
tion are considered coupled from grid-block to grid-block. Cao et al. [187] claimed
that light components cause more instability than heavy components, because light
components are more likely to stay in the gas phase, which has a much higher flow
rate than oil phase due to its small viscosity. So, if we include one extra variable x i ,
yi , or zi (component i is the lightest one) in the implicit variables, the new model
(IMPSAT++) could be much more stable than the IMPSAT model.
Hashem [189] used the information from previous time step to update the viscosity
in the transmissibility terms and viscosity didn’t update from previous iteration level
during iteration at certain time. Haukas et al. [188] showed that this leads to a better
convergence rate for the solution relative to solution techniques where viscosity
updated from previous iteration level during iteration at certain time [193].
Implicit Pressure Explicit Saturation and Composition (IMPESC). This tech-
nique was the first technique used to solve the compositional equations numerically
and also called IMPECS [175]. This method is less stable than other methods, because
only pressure is treated implicitly, but it is the fastest and cheapest method. Fussell
and Fussell [177] mentioned that explicit evaluation of composition might intro-
duce error in saturation pressure calculations. They suggested using smaller time
steps to overcome this error. Mansoori [194] discussed Nghiem et al. [181] method
248 A. Jamili et al.

and showed that their method would have fluctuation and slow convergence rate
especially when phase envelope crosses.
Sequential Implicit (SI). In this method, pressure and saturations are treated
implicitly and in sequential order. This method has been rarely used in the literature.
Watts [195] described the procedure as follows:
1. Construct a set of pressure equations and solve them for pressures at the new
time level.
2. Compute interblock total fluid velocities with these pressures.
3. Use interblock total velocities and the Buckly-Leverett phase velocity rela-
tionship to construct a set of saturation equations having, in general, implicit
treatment of relative permeabilities and capillary pressures. Solve these for
saturations at the new time level.
4. Use the saturations to compute interblock phase velocities.
5. Use the phase velocities to compute interblock component transport and the
amount of each component in each block at the end of the timestep.
6. Make the required fluid properties and function evaluation calculations and
proceed to the next timestep.

5.11 Case Study of Pressure Maintenance

Pressure maintenance has been proven to be one of the most widely and successful
methods for increasing recovery from a hydrocarbon reservoir. Injection of hydro-
carbon gas into the crest of the sensitivity analysis of the impact of geological uncer-
tainties on production forecasting (SAIGUP) fractured geometry was simulated by
Matlab Reservoir Simulation Toolbox (MRST). Dry gas will be injected through 3
injection wells into a reservoir contains volatile oil. The most important parameters
of a pressure maintenance project such as oil rate, average reservoir pressure and
GOR of the reservoir were investigated in order to estimate the success of pressure
maintenance in a naturally fractured reservoir.

5.11.1 Geometry Description

Geometry and rock properties were educed from SAIGUP project. SAIGUP is funded
by the European Commission (EC) [196]. The sedimentological model is a shoreface
sandstone reservoir were the deposition of sediments is due to the variation of the sea
level. The sedimentological model was made by studying 56 sedimentation systems
in different tectonically and climates situations. SAIGUP sedimentological model
stands as a heterogenetic and natural fractured model [197, 198].
The combination of sedimentological and structural model makes an overall
form which consists of three different fault system with variable fault density and
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 249

fault permeability characteristics. There is also 6 different rock type that each has
specific rock-fluid properties (relative permeability, porosity, net to gross ratio, capil-
lary function and etc.). The rock and fluid properties differences lead to reservoir
heterogeneity.
Geometry of the reservoir is structured from a standard Corner-Point method.
Hexahedral cells are put together based on a Cartesian model in a way that counting
them is easy. Fine-scale geometry contains 1.5 million cells of 40 × 120 × 20
dimensions and 37.5 × 37.5 × 1 cubic meter size. The upscaled geometry has
dimensions of 75 × 75 × 4 cubic meter and contains 96,000 cells [199]. The upscaling
method is based on warren method [200] but here, linear pressure drop was used
instead of No-Flow boundary condition. 78,720 active cells are separated from 96,000
cells by a No-Flow fault face on north side of the reservoir.

5.11.2 Rock Properties

Rock properties are related to a Cartesian geometry of 40 × 120 × 20 dimensions


based on geostatistical algorithm, then the same properties transferred to the main
Cartesian-point geometry. Figures 5.22 and 5.23 show the porosity on both Cartesian
and corner-point geometries. As there are 96,000 cells, this method helps to relate
all the rock properties to each cell easily.
Different rock properties such as porosity and permeability are related to each
cell, based on 6 different rock type. The fifth rock type has zero permeability and
almost zero porosity which stand as a barrier in the middle of the reservoir and

Fig. 5.22 Porosity on Cartesian geometry from angle view of −135° and height of 30
250 A. Jamili et al.

Fig. 5.23 Porosity on corner-point geometry from angle view of −65° and height of 55

Table 5.12 Rock properties


Parameter Average value Minimum Maximum
value value
Porosity 0.1166 0.0093 0.2911
Permeability in 126.7 0.0017 4101
x axis (md)
Permeability in 126.8 0.0018 4097
y axis (md)
Permeability in 14.7 0.0000 2172
z axis (md)

prevent fluid flow. This barrier effect is obvious in pressure distribution across the
reservoir. It is assumed that the reservoir contains high percentage of shale which
reduce permeability in cells with high concentration of shale. Mud also cover the
upper layers of reservoir which leads to decrease in lateral permeability. Average
rock porosity and permeability are shown in Table 5.12.

5.11.3 Fluid Properties

Fluid properties of volatile oil are extracted from the fifth comparative solution project
[201]. The compositional fluid description was a six component Peng-Robinson (PR)
characterization. The volatile oil contains the following mole percent: 50% C1 , 3%
C3 , 7% C6 , 20% C10 , 15% C15 , and 5% C20 . Peng-Robinson fluid description of
volatile oil is given in Table 5.13. All binary interaction coefficient are zero except:
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 251

Table 5.13 Peng-Robinson Fluid


Component Pc (psia) Tc (°R) MW Accen. Fac Critz
C1 667.8 343.0 16.040 0.0130 0.290
C3 616.3 665.7 44.100 0.1524 0.277
C6 436.9 913.4 86.180 0.3007 0.264
C10 304.0 1111.8 142.290 0.4885 0.257
C15 200.0 1270.0 206.000 0.6500 0.245
C20 162.0 1380.0 282.000 0.8500 0.235

Table 5.14 Reservoir data


Parameter Value
Water density (stock tank) 62.4 lb/ft3
Oil density (stock tank) 38.53 lb/ft3
Gas density (stock tank) 68.64 lb/MCF
Water compressibility 3.3 × 10–6 psi−1
Rock compressibility 5.0 × 10–6 psi−1
Water formation volume 1.000 RB/STB
factor
Water viscosity 0.7 cp
Reservoir temperature 160 °F
Oil formation volume factor −21.85 × 10–6 RB/STB/PSI
(slope above bubble point)
Initial water saturation 0.2
Initial oil saturation 0.8

C1 and C15 = 0.05, C1 and C20 = 0.05, C3 and C15 = 0.005 and C3 and C20 =
0.0 05.
Injected gas is a dry gas that consists of 94% of produced methane and 6% of
produced propane. Injection of produced gas is easy to gain, cheap and also prevent
the loss of reservoir energy. The initial reservoir saturation pressure is around 2300
psia and the range of minimum miscibility pressure is between 3000 to 3200 psia.
Reservoir data that used in this problem are given in Table 5.14.

5.11.4 Model Description

MATLAB Reservoir Simulation Toolbox (MRST) 2019a was used as simulator tool
in this study. MRST is an open-source software which uses a set of reservoir simu-
lators and workflow tools that can be modified in the way that suits the problem.
252 A. Jamili et al.

During simulation with MRST, we can use MATLAB/GNU octave to design specific
simulators or computational workflow [170].
Simulating a naturally fractured reservoir by a compositional model requires
solving complex and difficult linear and non-linear problems. In compositional
model, the phase behavior will be solved in each time-step and also we face changing
in unknowns as each phase may appear or vanish in each cell in every time step.
Automatic differentiation-object oriented (AD-OO) framework is used instead of
differentiation for reaching better solution. AD-OO framework solve the problem as
below:
1. Built a physical model where rock and fluid properties, well description and
more physical properties are defined properly.
2. Choosing the best non-linear solver for this specific problem.
3. Linearizing the system of non-linear problems.
4. Choosing the best linear solver.
Having 78,720 cells and also polarity of independent primary variables in this
problem makes us using automatic differentiation backend diagonal representation.
This backend AD has the advantage of compatibility with MATLAB executable (C++
Mex) which can be used for solving problems with number of independent primary
variables.
Generalized minimal residual method (GMRES) is the linear solver in this specific
compositional model. GMRES have the advantage of solving huge system of linear
equation and also can detect correct time step. Incomplete lower/upper (ILU) was
used as the one stage preconditioner [202]. Hierarchical fracture module (HFM) was
used for simulating fluid flow in MRST. This module is also known as embedded
discrete fracture model which is described in Sect. 5.10 [170].

5.11.5 Well Model

Inflow and outflow in MRST are in the subgrid scale and will be simulated by constant
bottom hole pressure (BHP) and or constant rate method. The main purpose of a
well model is to determine the pressure value around the well. Inflow-performance
relation (IPR) can determine relationship between pressure and rate around the well.
The Peaceman IPR relation was used in this simulating. 10 production and 3 injection
wells were designed for this problem. The injection wells are placed in the upper
part of the reservoir as shown in Fig. 5.24. All the wells operate under constant rate
with limiting BHP condition. More detail about wells is given in Table 5.15.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 253

Fig. 5.24 Upper view of wells and fault system

Table 5.15 Well description


Well type Rate BHP Radius of well Preferred fluid
Production 4000 STB/day 1000 psia 0.25 ft Water–Oil-Gas
Injection 5 MCF/day 4500 psia 0.25 ft Gas

5.11.6 Analysis of Results

Two scenario of production without any injection and production while pressure
maintenance starts from day 1 will be simulated and compared together. In the first
scenario, hydrocarbon produce by the use of natural reservoir energy which is a
sample of primary production. In this scenario, all the injection wells are closed
during the simulation, although the production wells are operated same as the second
scenario. In the second scenario, gas injection maintains reservoir pressure and a
higher recovery will be expected. Both scenarios are simulated for 15 years and then
the average pressure, GOR and oil rate of them will be compared below.
Average Pressure. First we investigate the results of average pressure to estimate
the effect of 15 years of gas injection on the reservoir energy. Higher reservoir average
pressure means the higher reservoir energy which leads to hydrocarbon production.
As seen in Fig. 5.25, gas injection maintains pressure of the reservoir. As average
pressure reaches the saturation pressure point, both gravity drainage and solution gas
drive mechanism activate which slow down the pressure drop. These mechanisms
almost have the same effect on both scenarios. However, the first scenario reaches
the saturation pressure sooner as there is no plan for maintaining pressure. At the
end of simulating time, the speed of pressure drop gradually decreases and average
pressure will be stable at the 1700 psia. This is the time for abounding the reservoir
or starting an EOR plan.
254 A. Jamili et al.

Fig. 5.25 Effect of gas injection on average reservoir pressure versus time for sc.1 and sc.2

As shown in Fig. 5.26, the pressure difference of two scenarios can be divided
into four periods: an initial production period when none of the scenarios reaches
the saturation pressure and pressure difference increases rapidly. At the end of this
period, the highest pressure difference value of 322 psia is recorded. As the first

Fig. 5.26 Pressure difference between two scenarios versus time (second scenario pressure—first
scenario pressure)
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 255

Fig. 5.27 GOR of the reservoir for both scenarios versus time

scenario reaches to saturation pressure and as the gravity drainage and solution gas
drive begin, the pressure difference tends to decreases. The final period results in
higher pressure in first scenario where the production is already stopped while in
second scenario, production at low rate is still continued.
Gas Oil Ratio. Figure 5.27 shows GOR during 15 years of simulating. Although,
production of the first scenario stops after 4395 days due to the low pressure level
but in the second scenario production last for 15 years. In the first scenario and
after 3633 days, production wells number 1 and 8 are the only open wells. And
after 4395 days, production well number 1 is shut down due to pressure drop and
production well number 8 is shut down because of high GOR above 10.
Almost till 3000 days, both scenarios have the same trend which means, injected
gas didn’t reach to the production wells. As the injected gas reaches the production
wells in the second scenario, GOR increases rapidly. In this specific problem, the
injected wells designed to reduce the possibility of gas canaling from the fractures.
Otherwise, the high GOR should happen at the early days of production which can
cause the production wells to shut down.
Oil Rate. As shown in Fig. 5.28, each scenario can be divided into three periods:
an initial production where all the production wells in both scenarios, produce with
constant rate of 4000 STB/Day condition. As the pressure of the reservoir drop
gradually, production wells are forced to produce under BHP condition. In the second
scenario, production well number 5 is the first well that produces under BHP condition
after 2745 Days, while the same well produces under BHP condition just after 1915
Days in the first scenario.
The first production well (number 9) will be shut down after 2466 and 3254 days,
respectively in scenarios 1 and 2. Respectively in scenario 1 and 2, after 3293 and
3879 days, only 4 wells are open to flow that can explain the third period. In the
256 A. Jamili et al.

Fig. 5.28 Oil production rate of all production wells versus time

first scenario, production wells number 1, 4, 6 and 8 will produce until the end with
respectively 1035, 1382, 815 and 1090 STB/Day.
Figure 5.29 represents the cumulative oil production in both scenarios. Almost
41 MMSTB more oil is produced from second scenario due to pressure maintenance
project. The production of oil in the second scenario will be continued till the last
day.

Fig. 5.29 Cumulative oil production during 15 years of simulating versus time
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 257

Appendices

Appendix A: Derivation of the Multiphase Flow Equations


in Compositional Simulation

The multiphase flow equations govern compositional simulation will be derived.


Basically, these equations are continuity equations for each component cover a
volume element xyz fixed in the space (Fig. 5.30) as following:

(Molar rate o f component c in) − (Molar rate o f component c out)


+ (Molar in jection rate o f component c)
− (Molar pr oduction rate o f component c)
= (Molar accumulation rate o f component c) (5.136)

Component c can be transported across the volume element boundary by two


mechanisms: diffusion and convection. These mechanisms, injection rate, production
rate, and accumulation rate will be discussed in details next.

Convection Mechanism

Convective transport is the amount of material carried along by the bulk movement
of the fluid. The driving force in convective transport is potential gradient. The molar
rate in minus molar rate out of component c (mole/time) for x, y, and z directions
due to convective transport in oil and gas phases are:
   
ρo xc vo,x x yz + ρg xc vg,x x yz
   
− ρo xc vo,x x+x yz − ρg xc vg,x x+x yz
   
+ ρo xc vo,y y xz + ρg xc vg,y y xz

Fig. 5.30 Volume element for deriving the multiphase flow equations
258 A. Jamili et al.
   
− ρo xc vo,y y+y xz − ρg xc vg,y y+y xz
   
+ ρo xc vo,z z xy + ρg xc vg,y z xy
   
− ρo xc vo,z z+z xy − ρg xc vg,z z+z xy = 0 (5.137)

where
kkr p −
→ → 

v p = − ∇ pp − γp ∇ D p = oil and gas (5.138)
μp

ρo and ρg are the molar densities of oil and gas.


φ is the porosity of the volume element.
xc is the mole fraction of component c in the oil phase.
yc is the mole fraction of component c in the gas phase.

Diffusion (Molecular) Transport

Molecular transport, or diffusion, can also add material across the faces of the volume
element. If No,c and Ng,c are the diffusion molar fluxes of component c (mole c per
time per area) in oil and gas phases, these quantities have units of mole per area per
time and represent the amount of transport by diffusion. Therefore, following the
previous approach, the molar rate in minus molar rate out of component c for x, y,
and z directions by diffusion are:
     
Nc,o,x x yz + Nc,g,x x yz − Nc,o,x x+x yz
     
− Nc,g,x x+x yz + Nc,o,y y xz + Nc,g,y y xz
   
− Nc,o,y y+y xz − Nc,g,y y+y xz
   
+ Nc,o,z z xy + Nc,g,z z xy
   
− Nc,o,z z+z xy − Nc,g,z z+z xy = 0 (5.139)

where
 
Nc,o = φρo So Dc,o ∇xc (5.140)

 
Nc,g = φρg Sg Dc,g ∇ yc (5.141)
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 259

Production or Injection

Finally, production and/or injection of component c into the volume element are
given by:

q D, f m,c + qC, f m,c (5.142)

where q D, f m and qC, f m are diffusion and convection mass transfer between matrix
and fracture at the matrix-fracture boundary.

Accumulation
 
 is φ ρo So + ρg Sg xyz,
The total mole of fluid in the volumeelement at any time
and the mole of component c is φ ρo So xc + ρg Sg yc xyz. Therefore, the rate
of accumulation of mole of component c is:

∂  
φ ρo So xc + ρg Sg yc xyz (5.143)
∂t

Flow Equations

The flow equations can be obtained by substituting Eqs. (5.137) to (5.143) into
Eq. (5.136). If the resulting equations are divided by volume element xyz and
applying limit when the volume element goes to zero, it becomes:
     
∂ ρo xc vo,x + ρg yc vg,x ∂ ρo xc vo,y + ρg yc vg,y ∂ ρo xc vo,z + ρg yc vg,z
− + +
∂x ∂y ∂z
     
∂ Nc,o,x + Nc,g,x ∂ Nc,o,y + Nc,g,y ∂ Nc,o,z + Nc,g,z
− + +
∂x ∂y ∂z
∂  
+ q D, f m,c + qC, f m,c = φ ρo So xc + ρg Sg yc (5.144)
∂t
Or, in vector notation,
   
− ∇. ρo xc vo + ρg yc vg − ∇. Nc,o + Nc,g + q D, f m,c + qC, f m,c
∂  
= φ ρo So xc + ρg Sg yc (5.145)
∂t
By substituting vo and vg from Eq. (5.138) and Nc,o and Nc,g from Eqs. (5.140)
and (5.141) into Eq. (5.145), it becomes:
260 A. Jamili et al.

kkr o −
→ → 
− kkrg −
→ → 

∇. ρo xc ∇ po − γo ∇ D + ρg yc ∇ pg − γg ∇ D
μo μg
    
+ ∇. φρo So Dc,o ∇xc + φρg Sg Dc,g ∇ yc + q D, f m,c + qC, f m,c
∂  
= φ ρo So xc + ρg Sg yc (5.146)
∂t
Equation (5.146) is a general case of compositional multiphase flow through
porous media for each component in oil and gas phases. For the water phase,
considering that hydrocarbon phases are immiscible in water, we have,

kkr w −
→ → 
− ∂
∇. ρw ∇ p w − γw ∇ D = [φ(ρw Sw )]. (5.147)
μw ∂t

Appendix B: Discretizing the Flow Equations

The final form of the general hydrocarbon flow equations, as obtained in Appendix
A (Eq. (5.147)), is as follows:
    n+1
∇ · ρo xc kkμroo ∇ p o − γo ∇
 D + ρg yc kkrg ∇  pg − γg ∇D
μg
    n+1 i, j,k
c = 1, 2, . . . , n c
+∇ · φρo So Dc,o ∇xc + φρg Sg Dc,g ∇ yc i, j,k
 n+1 ∂    n+1
+∇ · q D, f m,c + qC, f m,c i, j,k ∂t φ ρo So xc + ρg Sg yc i, j,k
(5.148)

For simplicity, flow equations are discretized in x-direction only. The same
procedure can be applied to y and z directions. Equation (5.148) in x-direction
becomes:
n+1
∂ kkr o ∂ po ∂D kkrg ∂ pg ∂D
ρo xc − γo + ρg yc − γg
∂x μo ∂t ∂t μg ∂t ∂t i, j,k
n+1
∂ ∂ xc ∂ yc
+ φρo So Dc,o + φρg Sg Dc,g
∂x ∂x ∂x i, j,k
 n+1 ∂    n+1
+ q D, f m,c + qC, f m,c i, j,k φ ρo So xc + ρg Sg yc i, j,k (5.149)
∂t

Discretization Oil and Gas Convective Terms in x-direction

Defining oil and gas convective terms as:


5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 261

k x k r o ∂ po ∂D
vox = ρo xc − γo (5.150)
μo ∂x ∂x

k x krg ∂ pg ∂D
vgx = ρg yc − γg (5.151)
μg ∂x ∂x

Oil convective flux term will be discretized first. By substituting Eq. (5.150) into
Eq. (5.149), the oil convective flux term in Eq. (5.149) in x-direction becomes:
$ %
∂ k x k r o ∂ po ∂ D n+1 ∂vo,x
n+1
ρo xc − γo =
i, j,k
(5.152)
∂x μo ∂x ∂x i, j,k ∂ x

Using central finite differences into Eq. (5.152):

∂vo,x
n+1 n+1
vo,x − vo,x
n+1
i+ 21 , j,k i− 21 , j,k
= xi, j,k = xi+ 21 , j,k − xi− 21 , j,k
i, j,k
(5.153)
∂x xi, j,k

where,
 n+1
vo,x n+11 = ρo xc kxμkor o
i+ 2 , j,k i+ 21 , j,k
$ %n+1 xi+ 21 , j,k = xi+1, j,k − xi, j,k
poi+1, j,k − poi, j,k Di+1, j,k −Di, j,k
x 1
− γoi+ 1 , j,k x 1
i+ 2 , j,k 2 i+ 2 , j,k

(5.154)

and,

k x kr o n+1
vo,x n+11 = ρo xc
i− 2 , j,k μo i− 1 , j,k
2
n+1
poi, j,k − poi−1, j,k Di, j,k − Di−1, j,k
− γoi− 1 , j,k (5.155)
xi− 21 , j,k 2 xi− 21 , j,k

By substituting Eqs. (5.154) and (5.155) into Eq. (5.152):

∂vg,
n+1
xi, j,k
∂x
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρ g x c μ g ⎠   ⎥
= ⎣ ( pgi+1, j,k − pgi, j,k ) − γg Di+1, j,k − Di, j,k ⎦
xi, j,k x i+ 21 , j,k
i+ 21 , j,k
262 A. Jamili et al.
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρ g x c μ g ⎠   ⎥
− ⎣ ( pgi, j,k − pgi−1, j,k ) − γg Di, j,k − Di−1, j,k ⎦
xi, j,k x i− 21 , j,k
i− 21 , j,k
(5.156)

Gas convective flux term can be discretized by the same manner as:

∂vg,x
n+1
i, j,k

∂x
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρ g x c μ g ⎠   ⎥
= ⎣ ( pgi+1, j,k − pgi, j,k ) − γg Di+1, j,k − Di, j,k ⎦
xi, j,k x i+ 21 , j,k
i+ 21 , j,k
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρ g x c μ g ⎠   ⎥
− ⎣ ( pgi, j,k − pgi−1, j,k ) − γg Di, j,k − Di−1, j,k ⎦
xi, j,k x i− 21 , j,k
i− 21 , j,k
(5.157)

Discretization the Oil and Gas Diffusive Flux Term in x-Direction

Oil diffusive flux term will be discretized first as:


n+1
∂ ∂ xc
φρo So Dc,o
∂x ∂x i, j,k
n+1 n+1 
1 ∂ xc ∂ xc
= φρo So Dc,o − φρo So Dc,o (5.158)
xi, j,k ∂x i+ 1 , j,k ∂x i− 1 , j,k
2 2

Expanding the (i+1/2) and (i−1/2) term in Eq. (5.158):


n+1
∂ xc φρo So Dc,o n+1  n+1
φρo So Dc,o = xci+1, j,k − xci, j,k (5.159)
∂x i+ 1 , j,k x i+ 1 , j,k
2 2

n+1 n+1
∂ xc φρo So Dc,o  n+1
φρo So Dc,o = xci, j,k − xci−1, j,k (5.160)
∂x i− 21 , j,k x i− 21 , j,k

Substituting Eqs. (5.159) and (5.160) in Eq. (5.158):

∂   n+1
φρo So Dc,o ∇xc i, j,k
∂x
1 φρo So Dc,o n+1  n+1
= xci+1, j,k − xci, j,k
xi, j,k x i+ 1 , j,k 2
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 263


φρo So Dc,o n+1  n+1
− xci, j,k − xci−1, j,k (5.161)
x i− 1 , j,k 2

The same procedure can be used to discretize gas molecular diffusion term as:

∂   n+1
φρg Sg Dc,g ∇ yc i, j,k
∂x
1 φρg Sg Dc,g n+1  n+1
= yci+1, j,k − yci, j,k
xi, j,k x i+ 21 , j,k
n+1 
φρg Sg Dc,g  n+1
− yci, j,k − yci−1, j,k (5.162)
x i− 1 , j,k 2

Discretization the Accumulation Term

The accumulation term is discretized by using regressive finite differences in time


as:
∂  
φ xc ρo So + yc ρg Sg i, j,k
∂t
1 . n+1  n /
= φxc ρo So + φyc ρg Sg − φxc ρo So + φyc ρg Sg (5.163)
t i, j,k

Final Form of Discretized Flow Equations

Substituting Eqs. (5.156), (5.157), (5.161), (5.162), and (5.163) into Eq. (5.149):
⎡ ⎤
k x kr o  n+1
1 ⎣ ρo x c μ o  

( poi+1, j,k − poi, j,k ) − γo Di+1, j,k − Di, j,k
xi, j,k x i+ 21 , j,k
i+ 21 , j,k
⎡ ⎤
k x kr o  n+1
1 ⎣ ρo x c μ o  
− ( poi, j,k − poi−1, j,k ) − γo Di, j,k − Di−1, j,k ⎦
xi, j,k x i− 21 , j,k
i− 21 , j,k
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρg yc μg ⎠   ⎥
+ ⎣ ( pgi+1, j,k − pgi, j,k ) − γg Di+1, j,k − Di, j,k ⎦
xi, j,k x i+ 21 , j,k
i+ 21 , j,k
⎡⎛ ⎞ ⎤n+1
k x krg
1 ⎢⎝ ρg yc μg ⎠   ⎥
− ⎣ ( pgi, j,k − pgi−1, j,k ) − γg Di, j,k − Di−1, j,k ⎦
xi, j,k x i− 21 , j,k
i− 21 , j,k
264 A. Jamili et al.

1 φρo So Dc,o n+1  n+1
+ xci+1, j,k − xci, j,k
xi, j,k x i+ 21 , j,k

φρo So Dc,o n+1  n+1
− xci, j,k − xci−1, j,k
x i− 21 , j,k

1 φρg Sg Dc,g n+1  n+1
+ yci+1, j,k − yci, j,k
xi, j,k x i+ 21 , j,k
n+1 
φρg Sg Dc,g  n+1
− yci, j,k − yci−1, j,k
x i− 21 , j,k
 n+1
+ q D, f m,c + qC, f m,c i, j,k
1 . n+1  n /
= φxc ρo So + φyc ρg Sg − φxc ρo So + φyc ρg Sg (5.164)
t i, j,k

Multiplying Eq. (5.164) by the volume of the grid cell, Vr,i, j,k =
yi, j,k z i, j,k xi, j,k , and rearranging, it becomes:
⎡  ⎤
yi, j,k z i, j,k ρo xc k xμkor o n+1
 
⎣ ( poi+1, j,k − poi, j,k ) − γo Di+1, j,k − Di, j,k ⎦
x i+ 21 , j,k
i+ 21 , j,k
⎡  ⎤
yi, j,k z i, j,k ρo xc k xμkor o n+1
 
−⎣ ( poi, j,k − poi−1, j,k ) − γo Di, j,k − Di−1, j,k ⎦
x i− 21 , j,k
i− 21 , j,k
⎡⎛ ⎞ ⎤n+1
k x krg
⎢ yi, j,k z i, j,k ρg yc μg   ⎥
+ ⎣⎝ ⎠ ( pgi+1, j,k − pgi, j,k ) − γg Di+1, j,k − Di, j,k ⎦
x i+ 21 , j,k
i+ 21 , j,k
⎡⎛ ⎞ ⎤n+1
k x krg
⎢ yi, j,k z i, j,k ρg yc μg   ⎥
− ⎣⎝ ⎠ ( pgi, j,k − pgi−1, j,k ) − γg Di, j,k − Di−1, j,k ⎦
x i− 21 , j,k
i− 21 , j,k

yi, j,k z i, j,k φρo So Dc,o n+1  n+1
+ xci+1, j,k − xci, j,k
x 1
i+ 2 , j,k
n+1 
yi, j,k z i, j,k φρo So Dc,o  n+1
− xci, j,k − xci−1, j,k
x i− 21 , j,k

φyi, j,k z i, j,k ρg Sg Dc,g n+1  n+1
+ yci+1, j,k − yci, j,k
x 1
i+ 2 , j,k
n+1 
yi, j,k z i, j,k φρg Sg Dc,g  n+1
− yci, j,k − yci−1, j,k
x i− 21 , j,k
 n+1
+ q D, f m,c + qC, f m,c i, j,k
Vr,i, j,k . n+1  n /
= φxc ρo So + φyc ρg Sg − φxc ρo So + φyc ρg Sg (5.165)
t i, j,k

Now, defining the transmissibility terms for oil and gas phases as:
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 265
 
yi, j,k z i, j,k ρo kxμkor o
Toi± 1 , j,k = (5.166)
2 x
i± 21 , j,k
⎛ ⎞
k x krg
yi, j,k z i, j,k ρg μg
Toi± 1 , j,k = ⎝ ⎠ (5.167)
2 x
i± 21 , j,k

yi, j,k z i, j,k φρo So Dc,o
M
To,c = (5.168)
i± 21 , j,k x i+ 21 , j,k

yi, j,k z i, j,k φρg Sg Dc,g
M
Tg,c = (5.169)
i± 1 , j,k
2 x i+ 21 , j,k

Substituting Eqs. (5.166) through (5.169) into Eq. (5.165), it becomes:


.   /n+1
(To xc )i+ 21 , j,k ( poi+1, j,k − poi, j,k ) − γoi+ 1 , j,k Di+1, j,k − Di, j,k
2
.   /n+1
− (To xc )i− 21 , j,k ( poi, j,k − poi−1, j,k ) − γoi− 1 , j,k Di, j,k − Di−1, j,k
2
.    /n+1
+ Tg yc i+ 1 , j,k ( pgi+1, j,k − pgi, j,k ) − γgi+ 1 , j,k Di+1, j,k − Di, j,k
2 2
.    /n+1
− Tg yc i− 1 , j,k ( pgi, j,k − pgi−1, j,k ) − γgi− 1 , j,k Di, j,k − Di−1, j,k
. 
2
 n+1  M n+1 
2
n+1 /
M n+1
+ To,c i+ 1 , j,k xci+1, j,k − xci, j,k − To,c i− 1 , j,k xci, j,k − xci−1, j,k
. 
2
 n+1  M n+1 
2
n+1 /
M n+1
+ Tg,c i+ 1 , j,k yci+1, j,k − yci, j,k − Tg,c i− 1 , j,k yci, j,k − yci−1, j,k
2 2
 n+1
+ q D, f m,c + qC, f m,c i, j,k
Vr,i, j,k . n+1  n /
= φxc ρo So + φyc ρg Sg − φxc ρo So + φyc ρg Sg (5.170)
t i, j,k

Equation (5.170) can also be written as:


 n+1     n+1
 To xc (po − (γo D)) i, j,k +  Tg yc pg −  γg D i, j,k
 M n+1  M n+1  n+1
+  To,c xc i, j,k +  Tg,c yc i, j,k + q D, f m,c + qC, f m,c i, j,k
Vr,i, j,k 
= t φxc ρo So + φyc ρg Sg (5.171)
t i, j,k

Equation (5.171) is the final discretized form of the hydrocarbon flow equations
in x direction. Following the same procedure, the discretized water flow equation in
x direction becomes:
 n+1 Vr,i, j,k
 Tw (pw − γw D) = t [φρw Sw ]i, j,k (5.172)
i, j,k t
266 A. Jamili et al.

where
 
yi, j,k z i, j,k ρw kxμkwr w
Twi± 1 , j,k = (5.173)
2 x
i± 21 , j,k

Appendix C: Newton–Raphson Method

The Newton–Raphson method to solve a set of nonlinear equations is described in


detail. The problem consists of solving the following set of non-linear equations:

F1 (x1 , x2 , . . . , xn )
F2 (x1 , x2 , . . . , xn )
·
(5.174)
·
·
Fn (x1 , x2 , . . . , xn )

or,

Fi (x) = 0 (5.175)

where F i , i = 1,2,…,n are the equations and x 1 ,x 2 , …,x n are the unknowns. To
develop the Newton–Raphson algorithm, all functions are first expressed as a Taylor
series expansion about an arbitrary point (x 1 ,x 2 , …,x n ,F 1 ,F 2 , …,F n ) as:

Fi (x1 + x1 , x2 + x2 , . . . , xn + xn )


∂ Fi ∂ Fi ∂ Fi
= Fi (x1 , x2 , . . . , xn ) + x1 + x2 + · · · + xn i = 1, 2, . . . , n
∂ x1 ∂ x2 ∂ xn
(5.176)

The objective is to find the roots of the equations by setting the left-hand sides of
these n equations equal to zero. If initial values of the unknowns are assumed, the
n equations of Eq. (5.157) can be solved for x1 , x2 ,…, xn . This system of n
equations may also be written as:
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 267

∂ F1 ∂ F1 ∂ F1
∂ x1
x1 + ∂ x2
x2 + ··· + ∂ xn
xn = −F1
∂ F2 ∂ F2 ∂ F2
∂ x1
x1 + ∂ x2
x2 + ··· + ∂ xn
xn = −F2
·
(5.177)
·
·
∂ Fn ∂ Fn ∂ Fn
∂ x1
x1 + ∂ x2
x2 + ··· + ∂ xn
xn = −F1

Equation (5.177) can also be expressed in matrix form as:


⎡ ∂ F1 ∂ F1 ⎤⎡ ⎤ ⎡ ⎤
∂ x1 ∂ x2
. . . ∂∂ Fxn1 x1 F1
⎢ ∂ F2 ∂ F2
. . . ∂∂ Fxn2 ⎥⎢ x ⎥ ⎢ F ⎥
⎢ ∂ x1 ∂ x2 ⎥⎢ 2 ⎥ ⎢ 2 ⎥
⎢ ⎥⎢ . ⎥ ⎢ . ⎥
⎢ . ⎥⎢ ⎥ ⎢ ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ . ⎥⎢ . ⎥ = ⎢ . ⎥ (5.178)
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎢ . ⎥⎢ . ⎥ ⎢ . ⎥
⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣ . ⎦⎣ . ⎦ ⎣ . ⎦
∂ Fn ∂ Fn
∂ x1 ∂ x2
. . . ∂∂ Fxnn xn Fn

Equation (5.178) can be written as:

[J ]x = −F (5.179)

J is called the Jacobian of the n equations system. The system of Eq. (5.179)
can be solved either by Gaussian elimination or by any appropriate procedure. The
unknowns (x 1 ,x 2 , …,x n ) are updated after each iteration as:

xil+1 = xil + xil+1 i = 1, 2, . . . , n (5.180)

where l is the iteration level. The iteration is terminated when max(|xi |) <
tolerance.

References

1. Terry RE. Enhanced oil recovery. Encycl Phys Sci Technol 2003; 18:503–18. https://doi.org/
10.1016/B0-12-227410-5/00868-1.
2. Klotz JA. The gravity drainage mechanism. J Pet Technol. 1953;5:19–21. https://doi.org/10.
2118/953019-G.
3. Jadhawar PS, Sarma HK. Numerical simulation and sensitivity analysis of gas-oil gravity
drainage process of enhanced oil recovery. J Can Pet Technol. 2010;49:64–70. https://doi.
org/10.2118/133373-PA.
4. Sandrea R, Nielsen RF. Dynamics of petroleum reservoirs under gas injection. Gulf Publishing
Company; 1974.
5. Buckley SE, Leverett MC. Mechanism of fluid displacement in sands. Trans AIME.
1942;146:107–16. https://doi.org/10.2118/942107-G.
268 A. Jamili et al.

6. Pirson SJ. Oil reservoir engineering. Inc, NY: Mc-Graw-Hill Book Co.; 1958.
7. Crawford PB. Pressure maintenance needed here. J Pet Technol. 1967;19:1449–52. https://
doi.org/10.2118/1784-PA.
8. Bennett E 0. Pressure maintenance. Drill. Prod. Pract., Amarillo, Texas: American Petroleum
Institute; 1938.
9. Tiwari S, Kumar MS. Nitrogen injection for simultaneous exploitation of gas cap. SPE Middle
East Oil Show, Society of Petroleum Engineers; 2001. https://doi.org/10.2118/68169-MS.
10. Gregory PP. Gas caps, their determination and significance. Trans AIME 1938; 127:25–30.
https://doi.org/10.2118/938025-G.
11. Bonett A, Pafitis D. Getting to the root of gas migration. Oilf Rev. 1996;8:36–49.
12. Riazi M, Alizadeh AH, Haghighi M, Sample R. Experimental study of gravity drainage during
gas injection in carbonate rocks. Middle East 2006; 1–7.
13. Craig FF, Geffen TM. The determination of partial pressure maintenance performance by
laboratory flow tests. Trans Am Inst Min Metall Eng. 1956;207:42–9.
14. Izadpanahi A, Azin R, Osfouri S, Malakooti R. Optimization of two simultaneous water and
gas injection scenarios in a high GOR Iranian oil field. In: 82nd EAGE Annu. Conf. Exhib.,
vol. 2020, European Association of Geoscientists & Engineers; 2020, p. 1–5.
15. Thomas F, Zhou X, Bennion D. Towards optimizing gas condensate reservoirs. Proc Annu
Tech Meet. 1995. https://doi.org/10.2118/95-09.
16. Kenyon D. Third SPE comparative solution project: gas cycling of retrograde condensate
reservoirs. J Pet Technol. 1987;39:981–97.
17. Crego WO, Henagan JM. Report on the Mamou field pressure maintenance project. J Pet
Technol. 1951;3:263–8. https://doi.org/10.2118/951263-G.
18. Benalycherif M, Bishlawi M, Bencheikh A. Pressure maintenance in the Zarzaitine Field-
Algeria 1972; 21.
19. Standing MB, Lindblad EN, Parsons RL. Calculated recoveries by cycling from a retrograde
reservoir of variable permeability. Trans AIME. 1948;174:165–90. https://doi.org/10.2118/
948165-G.
20. Spivak A, Dixon TN. Simulation of gas-condensate reservoirs. Proc SPE Symp Numer Simul
Reserv Perform 1973; 25. https://doi.org/10.2118/4271-MS.
21. Danesh A, Henderson GD, Krinis D, Peden JM. Experimental investigation of retrograde
condensation in porous media at reservoir conditions. In: SPE Annu. Tech. Conf. Exhib.,
Society of Petroleum Engineers; 1988. https://doi.org/10.2118/18316-MS.
22. Barbieri RR. Statistical secondary recovery model. SPE Adv Technol Ser. 1996;4:44–52.
https://doi.org/10.2118/26978-PA.
23. van Dijkum CE, Walker T. Fractured reservoir simulation and field development, Natih Field,
Oman. In: SPE Annu. Tech. Conf. Exhib., Society of Petroleum Engineers; 1991. https://doi.
org/10.2118/22917-MS.
24. Goss LE, Vague JR. Pressure maintenance operations-fort Chadbourne field odom lime reser-
voir. Fall Meet. Soc. Pet. Eng. AIME, Society of Petroleum Engineers; 1962. https://doi.org/
10.2118/406-MS.
25. Weyler JR, Sayre AT Jr. A novel pressure maintenance operation in a large stratigraphic trap.
J Pet Technol. 1959;11:13–7. https://doi.org/10.2118/1104-G.
26. Ma E, Gheorghiu S, Banagale M, Dashti L, Bond D, Ibrahim M, et al. Reservoir simulation
to support pressure maintenance projects in the greater Burgan Field, Kuwait. SPE Middle
East Oil Gas Show Conf., Society of Petroleum Engineers; 2015. https://doi.org/10.2118/172
517-MS.
27. Pavangat V, Patacchini L, Goyal P, Mohamed F, Lavenu APC, Aubertin F, et al. Development
of a giant carbonate oil field, part 1: fifty years of pressure maintenance history. Abu Dhabi
Int Pet Exhib Conf 2015; 1–20. https://doi.org/10.2118/177768-MS.
28. Smith GE. Waterflooding heavy oils. SPE Rocky Mt. Reg. Meet., society of petroleum
engineers; 1992, p. 473–80. https://doi.org/10.2118/24367-MS.
29. Pawelek J, Chorney M. Waterflood behaviour of the low-gravity wainwright pool. J Can Pet
Technol. 1966;5:97–104. https://doi.org/10.2118/66-02-08.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 269

30. Ezekwe N, Smith S, Wilson M, Humphrey M, Murphy J. Performance of a peripheral water-


flood project in the main body “B” reservoirs (stevens) at the elk hills oil field, California. SPE
West. Reg. Meet., Society of Petroleum Engineers; 1991. https://doi.org/10.2118/21759-MS.
31. Cason LD. Waterflooding increases gas recovery. J Pet Technol. 1989;41:1102–6. https://doi.
org/10.2118/12041-PA.
32. Beliveau D. Waterflooding viscous oil reservoirs. SPE Reserv Eval Eng. 2009;12:4–6. https://
doi.org/10.2118/113132-PA.
33. Woodling GS, Taylor PJ, Sun HH, Nguyen QN, Brix TR. Layered waterflood surveillance in
a mature field: the long beach unit. SPE West. Reg. Meet., Society of Petroleum Engineers;
1993. https://doi.org/10.2118/26082-MS.
34. Kasrale M, Sammon PH, Jespersen PJ. Field development options for a waterflooded heavy-oil
reservoir. J Pet Technol. 1993;45:888–94. https://doi.org/10.2118/20049-PA.
35. Oefelein FH, Walker JW. California flood yields profitable recovery of heavy oil from
multilayered reservoir. J Pet Technol. 1964;16:509–14. https://doi.org/10.2118/699-PA.
36. Moghadasi J, Jamialahmadi M, Müller-Steinhagen H, Sharif A, Izadpanah MR, Motaei E,
et al. Formation damage in Iranian oil fields. Int. Symp. Exhib. Form. Damage Control: Society
of Petroleum Engineers; 2002.
37. Harley YJ. Waterflood behavior of high viscosity crudes in preserved soft and unconsolidated
cores. J Pet Technol. 1966; 18. https://doi.org/10.2118/1202-PA.
38. Roark GR. Analysis of waterflood performance–block VI Ranger Zone, Wilmington field.
Calif J Pet Technol. 1960;12:45–51. https://doi.org/10.2118/1370-G.
39. Nelson KC. Wilmington Townlot unit—waterflood implementation and response. In: SPE
Annu. Fall Tech. Conf. Exhib., Society of Petroleum Engineers; 1976. https://doi.org/10.
2118/6167-MS.
40. Jenkins C, Al-Sharif S, Harris R, Weisgram J, Michel D. Forty years of improved oil recovery:
lessons from low-permeability turbidites of the East Wilmington Field. California SPE Int
Pet Conf. 2004. https://doi.org/10.2118/92036-MS.
41. Adams D. Experiences with waterflooding lloydminster heavy-oil reservoirs. J Can Pet
Technol 1982:1643–50. https://doi.org/10.2118/10196-PA.
42. Edgson JJ, Czyzewski G. Production evaluation of the viking Kinsella B pool unit no. 1. J
Can Pet Technol 1985; 24:70–6.
43. Lim FH, Saner WB, Stilwell WH, Patton JT. Steamflood pilot test in waterflooded, 2,500-ft
Tar Zone reservoir, fault block II Unit, Wilmington Field, California. In: SPE Annu. Tech.
Conf. Exhib., Society of Petroleum Engineers; 1993, p. 1–16. https://doi.org/10.2118/266
15-MS.
44. Luis RJ, Navia W. Natural waterflooding, Campo La Peña. SPE J; 2001.
45. Dorsey JB. Performance review of shoats creek unit vaporizing gas-drive project. J Pet
Technol. 1968;20:416–22. https://doi.org/10.2118/1886-PA.
46. McGraw JH, Lohec RE. The Pickton field–review of a successful gas injection project. J Pet
Technol. 1964;16:399–405. https://doi.org/10.2118/667-PA.
47. Hermansen H, Thomas LKK, Sylte JEE, Aasboe BTT. Twenty five years of Ekofisk reservoir
management. SPE Annu Tech Conf Exhib. 1997. https://doi.org/10.2118/38927-MS.
48. Kaye E. Some factors in the economics of recycling. Trans AIME. 1942;146:221–30. https://
doi.org/10.2118/942221-G.
49. Brickner JW. Jurassic smackover development and pressure maintenance Bryans Mill Field,
Texas. In: Fall Meet. Soc. Pet. Eng. AIME, Society of Petroleum Engineers; 1965. https://
doi.org/10.2118/1278-MS.
50. Tarner J, Evans WR, Kaveler HH, Aime M, Co PP. The Shuler Jones sand pool; Nine years
of unitized pressure-maintenance operations 1951; 192:121–6.
51. Jacoby RH, Berry VJ. A method for predicting pressure maintenance performance for
reservoirs producing volatile crude oil. Pet Trans. 1958;213:59–64.
52. Wikipedia. Simple English Wikipedia the free Encyclopedia; 2010.
53. Morales E, Lee WJ. Lease fuel, fuel gas or fuel consumed in operations—is it reserves? In:
SPE Annu. Tech. Conf. Exhib., Society of Petroleum Engineers; 2014, p. 27–9. https://doi.
org/10.2118/170612-MS.
270 A. Jamili et al.

54. Zhang F, Fan X, Ding J. Field experiment of enhancing heavy oil recovery by cyclic fuel gas
injection. SPE J; 2000.
55. Barstow WF, Watt GW. Fifteen years of progress in catalytic treating of exhaust gas. In: SPE
Rocky Mt. Reg. Meet., Society of Petroleum Engineers; 1975. https://doi.org/10.2118/534
7-MS.
56. Kirtley JR. Nitrogen from cryogenic air separation process to be used for pressure maintenance
and to enhance recovery at Cantarell complex in Campeche Bay, Mexico. In: Offshore Technol.
Conf., Offshore Technology Conference; 1999. https://doi.org/10.4043/10864-MS.
57. Stewart FM, Garthwaite DL, Karebill FK. Pressure maintenance by inert gas injection in the
high Relife Elk Basin Field 1955; 204.
58. Thomas L, Dixon TN, Pierson RG, Hermansen H. Ekofisk nitrogen injection SPE Form Eval.
1991;6:151–60. https://doi.org/10.2118/19839-PA.
59. Zaman K, el Moemen MA. Energy consumption, carbon dioxide emissions and economic
development: evaluating alternative and plausible environmental hypothesis for sustainable
growth. Renew Sustain Energy Rev. 2017;74:1119–30. https://doi.org/10.1016/j.rser.2017.
02.072.
60. Benson SM, Orr FM Jr. Carbon dioxide capture and storage. MRS Bull. 2008;33:303–5.
https://doi.org/10.1557/mrs2008.63.
61. Law DH-S. Disposal of carbon dioxide, a greenhouse gas, for pressure maintenance in a
steam-based thermal process for recovery of heavy oil and bitumen. SPE Int. Therm. Oper.
Heavy Oil Symp. West. Reg. Meet., Society of Petroleum Engineers; 2004. https://doi.org/
10.2118/86958-MS.
62. Zheng S, Li H, Yang D. Pressure maintenance and improving oil recovery with CO2 injection
in heavy oil reservoirs. Wire 2011; 2.
63. Zheng S, Li H, Yang D. Pressure maintenance and improving oil recovery with immiscible
CO2 injection in thin heavy oil reservoirs. J Pet Sci Eng. 2013;112:139–52.
64. Pourhadi S, Fath AH. Performance of the injection of different gases for enhanced oil recovery
in a compositionally grading oil reservoir. J Pet Explor Prod Technol. 2020;10:641–61.
65. Bird RB, Stewart WE, Lightfoot EN, Meredith RE. Transport phenomena. J Electrochem Soc.
1961;108:78C.
66. Cussler EL, Cussler EL. Diffusion: mass transfer in fluid systems. Cambridge university press;
2009.
67. Chukwuma FO. Mass transfer of gaseous carbon dioxide into a quiescent liquid hydrocarbon.
University of Tulsa [Oklahoma]; 1983.
68. Bird RB, Stewart WE, Lightfoot EN. Transport phenomena, 2nd edition (2002) n.d.
69. Da Silva FV, Belery P. Molecular diffusion in naturally fractured reservoirs: a decisive recovery
mechanism. SPE Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 1989.
70. Sigmund PM. Prediction of molecular diffusion at reservoir conditions. Part II-estimating the
effects of molecular diffusion and convective mixing in multicomponent systems. J Can Pet
Technol 1976; 15.
71. Unver AA, Himmelblau DM. Diffusion coefficients of CO2 , C2 H4 , C3 H6 and C4 H8 in Water
from 6 to 65 C. J Chem Eng Data. 1964;9:428–31.
72. Thomas WJ, Adams MJ. Measurement of the diffusion coefficients of carbon dioxide and
nitrous oxide in water and aqueous solutions of glycerol. Trans Faraday Soc. 1965;61:668–73.
73. Grogan AT, Pinczewski WV. The role of molecular diffusion processes in tertiary CO2
flooding. J Pet Technol. 1987;39:591–602.
74. Campbell BT, Orr FM Jr. Flow visualization for CO2 /crude-oil displacements. Soc Pet Eng
J. 1985;25:665–78.
75. Amyx J, Bass D, Whiting RL. Petroleum reservoir engineering physical properties; 1960.
76. Langnes GL, Robertson Jr JO, Chilingar G V. Secondary recovery and carbonate reservoirs;
1972.
77. Stewart FM, Garthwaite DL, Karebill FK. Pressure maintenance by inert gas injection in the
high relief elk basin field 1955; 204.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 271

78. Kelly P, Kennedy SL. Thirty years of effective pressure maintenance by gas injection in the
Hilbig field. J Pet Technol. 1965;17:279–81. https://doi.org/10.2118/946-PA.
79. Cotter WH. Twenty-three years of gas injection into a highly undersaturated crude reservoir.
J Pet Technol. 1962;14:361–5.
80. Wei MH, Yu JP, Moore DM, Ezekwe N, Querin ME, Williams LL. Case history of pressure
maintenance by Crestal gas injection in the 26R gravity drainage reservoir. In: SPE West.
Reg. Meet., Society of Petroleum Engineers; 1992. https://doi.org/10.2118/24035-MS.
81. van Wingen N, Barton CW, Case CH. Coalinga nose pressure maintenance project. J Pet
Technol. 1973;25:1147–52. https://doi.org/10.2118/4183-PA.
82. Waller RL. The Northwest Avard pressure maintenance project. J Pet Technol. 1971;23:1074–
80. https://doi.org/10.2118/3156-PA.
83. Babson EC. A review of gas injection projects in California. SPE Calif. Reg. Meet.: Society
of Petroleum Engineers; 1989.
84. McQuillan H. Fracture-controlled production from the Oligo-Miocene Asmari Formation in
Gachsaran and Bibi Hakimeh fields, southwest Iran. Carbonate Pet. Reserv.: Springer; 1985.
p. 511–23.
85. Saidi AM. Twenty years of gas injection history into well-fractured Haft Kel Field (Iran). In:
Int. Pet. Conf. Exhib. Mex., Society of Petroleum Engineers; 1996. https://doi.org/10.2118/
35309-MS.
86. Vowel WR. Gas injection in Eastern Venezuela field increases oil recovery. J Pet Technol.
1958;10:31–3.
87. Schilthuis RJ. Active oil and reservoir energy. Trans AIME. 1936;118:33–52. https://doi.org/
10.2118/936033-G.
88. Muskat M. Physical principles of oil production. McGraw-Hill Book Co.; 1949.
89. Havlena D, Odeh AS. The material balance as an equation of a straight line. J Pet Technol.
1963;15:896–900. https://doi.org/10.2118/559-PA.
90. Sandrea R, Nielsen RF. Combination drive predictions by the Muskat and differential tracy
material balances using various empirical relations and theoretical saturation equations. J Can
Pet Technol. 1969;8:98–104. https://doi.org/10.2118/69-03-04.
91. van Golf-Racht TD. Fundamentals of fractured reservoir engineering. vol. 12. Elsevier; 1982.
92. Hinkley RE, Davis LA. Capillary pressure discontinuities and end effects in homogeneous
composite cores: effect of flow rate and wettability. SPE Annu. Tech. Conf. Exhib.: Society
of Petroleum Engineers; 1986.
93. Willhite GP. Waterflooding, vol. 3. Richardson, Texas Textb Ser SPE; 1986.
94. Dindoruk B, Firoozabadi A. Computation of gas-liquid drainage in fractured porous media
recognizing fracture liquid flow. Annu. Tech. Meet.: Petroleum Society of Canada; 1994.
95. Firoozabadi A, Hauge J. Capillary pressure in fractured porous media (includes associated
papers 21892 and 22212). J Pet Technol. 1990;42:784–91.
96. Rangel-German ER, Kovscek AR. Microvisual analysis of matrix-fracture interaction. SPE
Int. Pet. Conf. Mex.: Society of Petroleum Engineers; 2004.
97. Witherspoon PA, Wang JSY, Iwai K, Gale JE. Validity of cubic law for fluid flow in a
deformable rock fracture. Water Resour Res. 1980;16:1016–24.
98. McDonald AE, Beckner BL, Chan HM, Jones TA, Wooten SO. Some important considerations
in the simulation of naturally fractured reservoirs. Low Permeability Reserv. Symp.: Society
of Petroleum Engineers; 1991.
99. Jones TA, Wooten SO, Kaluza TJ. Single-phase flow through natural fractures. SPE Annu.
Tech. Conf. Exhib.: Society of Petroleum Engineers; 1988.
100. Saidi AM, Tehrani DH, Wit K. PD 10 (3) mathematical simulation of fractured reservoir
performance, based on physical model experiments. 10th World Pet. Congr., World Petroleum
Congress; 1979.
101. Horie T, Firoozabadi A, Ishimoto K. Laboratory studies of capillary interaction in frac-
ture/matrix systems. SPE Reserv Eng. 1990;5:353–60.
102. Saidi AM. Discussion of capillary-pressure in fractured porous-media. J Pet Technol.
1991;43:233–5.
272 A. Jamili et al.

103. Saidi AM. Reservoir engineering of fractured reservoirs (fundamental and Practical Aspects).
Total; 1987.
104. Sajadian VA, Danesh A, Tehrani DH. Laboratory studies of gravity drainage mechanism in
fractured carbonate reservoir-capillary continuity. Abu Dhabi Int. Pet. Exhib. Conf.: Society
of Petroleum Engineers; 1998.
105. Firoozabadi A, Ottesen B, Mikklesen M. Measurements of supersaturation and critical gas
saturation (includes associated papers 27920 and 28669). SPE Form Eval. 1992;7:337–44.
https://doi.org/10.2118/19694-PA.
106. Kazemi H, Merrill LS Jr, Porterfield KL, Zeman PR. Numerical simulation of water-oil flow
in naturally fractured reservoirs. Soc Pet Eng J. 1976;16:317–26.
107. Saidi AM, Sakthikumar S. Gas gravity drainage under secondary and tertiary conditions in
fractured reservoirs. Middle East Oil Show: Society of Petroleum Engineers; 1993.
108. Labastie A. Capillary continuity between blocks of a fractured reservoir. SPE Annu. Tech.
Conf. Exhib.: Society of Petroleum Engineers; 1990.
109. Festoy S, Golf-Racht V. Gas gravity drainage in fractured reservoirs through new dual-
continuum approach (includes associated papers 20296 and 20390). SPE Reserv Eng.
1989;4:271–8.
110. Thomas LK, Dixon TN, Pierson RG. Fractured reservoir simulation SPE J. 1983;23:42–54.
https://doi.org/10.2118/9305-PA.
111. Fung LSK, Collins DA. An evaluation of the improved dual porosity model for the simulation
of gravity effects in naturally fractured reservoirs. J Can Pet Technol 1991; 30.
112. Coats KH, Nielsen RL, Terhune MH, Weber AG. Simulation of three-dimensional, two-phase
flow in oil and gas reservoirs. Soc Pet Eng J. 1967;7:377–88.
113. Boerrigter P, Maas JG, de Vries A. A fractured reservoir simulator capable of modeling
block-block interaction. SPE Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers;
1989.
114. Coats KH. Implicit compositional simulation of single-porosity and dual-porosity reservoirs.
SPE Symp. Reserv. Simul.: Society of Petroleum Engineers; 1989.
115. Vidal J. Gas/oil gravity drainage: laboratory study of oil reimbibition and streaming between
stacked matrix blocks. Eur. Pet. Conf.: Society of Petroleum Engineers; 1992.
116. Cardwell WT Jr, Parsons RL. Gravity drainage theory Trans AIME. 1949;179:199–215.
https://doi.org/10.2118/949199-G.
117. Firoozabadi A, Ishimoto K, Dindoruk B. Theory of reinfiltration in fractured porous media:
part I-one-dimensional model. Pap SPE. 1991;21796:20–2.
118. Tan JCT, Firoozabadi A. Dual-porosity simulation incorporating reinfiltration and capillary
continuity concepts part i: single gridcell. SPE Reserv. Simul. Symp.: Society of Petroleum
Engineers; 1995.
119. Horie T, Firoozabadi A, Ishimoto K. Capillary continuity in fractured reservoirs SPE.
1988;18282:2–5.
120. Le Gallo Y, Le Romancer JF, Bourbiaux B, Fernandes G. Mass transfer in fractured reservoirs
during gas injection: experimental and numerical modeling. SPE Annu. Tech. Conf. Exhib.:
Society of Petroleum Engineers; 1997.
121. Qasem FH, Ershaghi I. Optimizing gas injection into naturally fractured reservoirs. SPE Lat.
Am. Pet. Eng. Conf.: Society of Petroleum Engineers; 1994.
122. Eica AS, Ershaghi I. Modeling of gas injection in fractured reservoirs. SPE Calif. Reg. Meet.:
Society of Petroleum Engineers; 1982.
123. Yamamoto RH, Padgett JB, Ford WT, Boubeguira A. Compositional reservoir simulator for
fissured systems—the single-block model. Soc Pet Eng J. 1971;11:113–28. https://doi.org/
10.2118/2666-PA.
124. Dykstra H. The prediction of oil recovery by gravity drainage. J Pet Technol. 1978;30:818–30.
125. Schechter DS, Guo B. Mathematical modeling of gravity drainage after gas injection into
fractured reservoirs. Permian Basin Oil Gas Recover. Conf.: Society of Petroleum Engineers;
1996.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 273

126. Nenniger E Jr, Storrow JA. Drainage of packed beds in gravitational and centrifugal-force
fields. AIChE J. 1958;4:305–16.
127. Pavone D, Bruzzi P, Verre R. Gravity drainage at low IFT. Gravity Drain. low IFT. In: 5th
Eur. Symp. Enhanc. Oil Recover. Budapest; 1989.
128. Luan Z-A. Some theoretical aspects of gravity drainage in naturally fractured reservoirs. SPE
Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 1994.
129. Aronofsky JS, Masse L, Natanson SG. A model for the mechanism of oil recovery from the
porous matrix due to water invasion in fractured reservoirs. Trans AIME. 1958;213:17–9.
130. Martin JC. Reservoir analysis for pressure maintenance operations based on complete
segregation of mobile fluids v v ; 1958.
131. Maryam H, Reza A, Rouhollah F, Sohrab Z. New insights into forced and free fall gravity
drainage performance in a fractured physical model. J Pet Sci Eng. 2021.
132. Morel DD, Bourbiaux B, Latil M, Thiebot B. diffusion effects in gas-flooded light oil fractured
reservoir. Paper SPE 20516 presented at the 65th Annual Technical Conf. and Exh. New
Orleans; 1990. p. 23–6.
133. Renner TA. Measurement and correlation of diffusion coefficients for CO2 and rich-gas
applications. SPE Reserv Eng. 1988;3:517–23.
134. Thiebot BM, Sakthikumar SS. Cycling fractured reservoirs containing volatile oil: laboratory
investigation of the performance of lean gas or nitrogen injection. Middle East Oil Show:
Society of Petroleum Engineers; 1991.
135. Le Romancer JF, Defives D, Kalaydjian F, Fernandes G. Influence of the diffusing gas on the
mechanism of oil recovery by gas diffusion in fractured reservoir. IEA Collab. Proj. Enhanc.
Oil Recover. Work. Symp. Bergen, Norw.; 1994. p. 28–31.
136. Le Romancer JF, Fernandes G. Mechanism of oil recovery by gas diffusion in fractured
reservoir in presence of water. SPE/DOE Improv. Oil Recover. Symp.: Society of Petroleum
Engineers; 1994.
137. Riazi MR, Whitson CH, Da Silva F. Modelling of diffusional mass transfer in naturally
fractured reservoirs. J Pet Sci Eng. 1994;10:239–53.
138. Darvish GR, Lindeberg EGB, Holt T, Kleppe J, Utne SA. Reservoir conditions laboratory
experiments of CO2 injection into fractured cores. SPE Eur. Annu. Conf. Exhib.: Society of
Petroleum Engineers; 2006.
139. Karimaie H. Aspects of water and gas injection in fractured reservoirs; 2007.
140. Thomas LK, Dixon TN, Pierson RG, Hermansen H. Ekofisk nitrogen injection. SPE Form
Eval. 1991; 6:151–60. https://doi.org/10.2118/19839-PA.
141. Hu H, Whitson CH, Qi Y. A study of recovery mechanisms in a nitrogen diffusion experiment.
In: IOR 1991–6th Eur. Symp. Improv. Oil Recover., European Association of Geoscientists
& Engineers; 1991. p. cp-44.
142. Fayers FJ, Lee S-T. Crossflow mechanisms by gas drive in hetergeneous reservoirs. SPE
Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 1992.
143. Lenormand R, Le Romancer JF, Le Gallo Y, Bourbiaux B. Modeling the diffusion flux between
matrix and fissure in a fissured reservoir. SPE Annu. Tech. Conf. Exhib.: Society of Petroleum
Engineers; 1998.
144. Hoteit H, Firoozabadi A. Numerical modeling of diffusion in fractured media for gas injection
and recycling schemes. SPE Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 2006.
145. Alavian SA, Whitson CH. Modeling CO2 injection in a fractured chalk experiment.
In: SPE/EAGE Reserv. Charact. Simul. Conf., European Association of Geoscientists &
Engineers; 2009. p. cp-170.
146. Moortgat J, Firoozabadi A, Farshi MM. A new approach to compositional modeling of CO2
injection in fractured media compared to experimental data. SPE Annu. Tech. Conf. Exhib.:
Society of Petroleum Engineers; 2009.
147. Cussler EL. Multicomponent diffusion. Elsevier; 2013.
148. Firoozabadi A, Ghorayeb K, Shukla K. Theoretical model of thermal diffusion factors in
multicomponent mixtures. AIChE J. 2000;46:892–900.
274 A. Jamili et al.

149. Sigmund PM. Prediction of molecular diffusion at reservoir conditions. Part 1-Measurement
and prediction of binary dense gas diffusion coefficients. J Can Pet Technol 1976; 15.
150. Charbeneau RJ. Groundwater hydraulics and pollutant transport. Waveland Press; 2006.
151. Green DW, Willhite GP. Enhanced oil recovery, vol. 6. Henry L. Doherty Memorial Fund of
AIME: Society of Petroleum Engineers Richardson, TX; 1998.
152. Lagalaye Y, Nectoux A, James N. Characterization of acid gas diffusion in a carbonate frac-
tured reservoir through experimental studies, numerical simulation and field pilots. SPE Annu.
Tech. Conf. Exhib.: Society of Petroleum Engineers; 2002.
153. Ramirez BA, Kazemi H, Al-Kobaisi M, Ozkan E, Atan S. A critical review for proper use
of water-oil-gas transfer functions in dual-porosity naturally fractured reservoirs-part I. SPE
Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 2007.
154. Kimbler OK, Caudle BH. Areal sweepout behavior in a nine-spot injection pattern. J Pet
Technol. 1964;16:199–202. https://doi.org/10.2118/784-PA.
155. Dyes AB, Caudle BH, Erickson RA. Oil production after breakthrough as influenced by
mobility ratio. J Pet Technol. 1954;6:27–32. https://doi.org/10.2118/309-G.
156. Aronofsky JS, Ramey HJ. Mobility ratio—its influence on injection or production histories
in five-spot water flood. J Pet Technol. 1956;8:205–10. https://doi.org/10.2118/641-G.
157. Aronofsky JS. Mobility ratio—its influence on flood patterns during water encroachment. J
Pet Technol. 1952;4:15–24. https://doi.org/10.2118/132-G.
158. Craig, Geffen TM, Morse RA, Craig Jr. FF, Geffen TM, Morse RA. Oil recovery performance
of pattern gas or water injection operations from model tests. Pet Branch Fall Meet San
Antonio. 1955; 204:1–9.
159. Collings RC, Hild GP, Abidi HR. Pattern modification by injection-well shut-in: a combined
cost-reduction and sweep-improvement effort. SPE Reserv Eng. 1996;11:69–72. https://doi.
org/10.2118/30730-PA.
160. Moradi B, Tangsiri Fard J, Rasaei MR, Momeni A, Bagheri MB. Effect of gas recycling on
the enhancement of condensate recovery in an Iranian fractured gas/condensate reservoir.
Trinidad Tobago Energy Resour. Conf., Society of Petroleum Engineers; 2010. https://doi.
org/10.2118/132840-MS.
161. Dixon PC, Boatright BB. Practical economics of cycling; 1941.
162. Prats HF. The application of numerical methods to cycling and flooding problems. 3rd World
Pet Congr; 1951.
163. Latil M. Enhanced oil recovery. Éditions Technip; 1980.
164. Fung L-K. Numerical simulation of naturally fractured reservoirs. Middle East Oil Show:
Society of Petroleum Engineers; 1993.
165. Odeh AS. Comparison of solutions to a three-dimensional black-oil reservoir simulation
problem (includes associated paper 9741). J Pet Technol. 1981;33:13–25. https://doi.org/10.
2118/9723-PA.
166. Behie A, Vinsome PKW. Block iterative methods for fully implicit reservoir simulation. Soc
Pet Eng J 1982; 22:658–68. https://doi.org/10.2118/9303-PA.
167. Bansal PP, McDonald AE, Moreland EE, Trimble RH. A strongly coupled, fully implicit, three
dimensional, three phase reservoir simulator. In: SPE Annu. Tech. Conf. Exhib., Society of
Petroleum Engineers; 1979. https://doi.org/10.2118/8329-MS.
168. Kazemi H, Gilman JR, Elsharkawy AM. Analytical and numerical solution of oil recovery
from fractured reservoirs with empirical transfer functions (includes associated papers 25528
and 25818). SPE Reserv Eng. 1992;7:219–27.
169. Pruess K, Narasimhan TN. A practical method for modeling fluid and heat flow in fractured
porous media; 1982.
170. Lie K-A. An introduction to reservoir simulation using MATLAB; 2016.
171. Shah S, Møyner O, Tene M, Lie K-A, Hajibeygi H. The multiscale restriction smoothed basis
method for fractured porous media (F-MsRSB). J Comput Phys. 2016;318:36–57. https://doi.
org/10.1016/j.jcp.2016.05.001.
172. Monteagudo JEP, Firoozabadi A. Control-volume method for numerical simulation of two-
phase immiscible flow in two-and three-dimensional discrete-fractured media. Water Resour
Res. 2004; 40.
5 Gas Injection for Pressure Maintenance in Fractured Reservoirs 275

173. Hoteit H, Firoozabadi A. Multicomponent fluid flow by discontinuous Galerkin and mixed
methods in unfractured and fractured media. Water Resour Res. 2005; 41.
174. Lake LW, Carey GF, Pope GA, Sepehrnoori K. Isothermal, multiphase, multicomponent fluid
flow in permeable media. In Situ;(United States) 1984; 8.
175. Young LC, Stephenson RE. A generalized compositional approach for reservoir simulation.
Soc Pet Eng J. 1983;23:727–42.
176. Coats KH. SPE 8284 an equation of state compositional model; 1980.
177. Fussell LT, Fussell DD. An iterative technique for compositional reservoir models. Soc Pet
Eng J. 1979;19:211–20. https://doi.org/10.2118/6891-PA.
178. Roebuck IF Jr, Henderson GE, Douglas J Jr, Ford WT. The compositional reservoir simulator:
case I-the linear model. Soc Pet Eng J. 1969;9:115–30.
179. Roebuck IF Jr, Ford WT, Henderson GE, Douglas J Jr. The compositional reservoir simulator.
Case IV The Two-Dimensional Model. Fall Meet. Soc. Pet. Eng. AIME: Society of Petroleum
Engineers; 1968.
180. Kazemi H, Vestal CR, Shank DG. An efficient multicomponent numerical simulator. Soc Pet
Eng J. 1978;18:355–68.
181. Nghiem LX, Fong DK, Aziz K. Compositional modeling with an equation of state (includes
associated papers 10894 and 10903). Soc Pet Eng J. 1981;21:687–98.
182. Coats KH, Thomas LK, Pierson RG. Compositional and black oil reservoir simulation. SPE
Reserv. Simul. Symp.: Society of Petroleum Engineers; 1995.
183. Coats KH, Thomas LK, Pierson RG. Compositional and black oil reservoir simulation. SPE
Reserv Eval Eng. 1998;1:372–9. https://doi.org/10.2118/50990-PA.
184. Wang P, Yotov I, Wheeler M, Arbogast T, Dawson C, Parashar M, et al. A new generation
EOS compositional reservoir simulator: Part I-formulation and discretization. SPE Reserv.
Simul. Symp.: Society of Petroleum Engineers; 1997.
185. Buchwalter JL. A new simplified compositional simulator; 1994.
186. Thiele KJ, Lake LW, Sepehrnoori KA. A comparison of three equation of state compositional
simulator. In: Proceeding 7th SPE Symp. Reserv. Simulation, San Fr. CA, Nov;1983. p. 16–8.
187. Cao H, Aziz K. Performance of IMPSAT and IMPSAT-AIM models in compositional
simulation. SPE Annu. Tech. Conf. Exhib.: Society of Petroleum Engineers; 2002.
188. Haukas J, Aavatsmark I, Espedal M, Reiso E. A comparison of two different impsat models
in compositional simulation. SPE J. 2007;12:145–51.
189. Hashem YKS. Explicit composition implicit pressure and saturation simulation of dual-
porosity/permeability reservoirs; 1999.
190. Quandalle P, Savary D. An implicit in pressure and saturations approach to fully compositional
simulation. SPE Symp. Reserv. Simul.: Society of Petroleum Engineers; 1989.
191. Rodriguez F, Galindo-Nava A. A general formulation for compositional reservoir simulation.
Int. Pet. Conf. Exhib. Mex.: Society of Petroleum Engineers; 1994.
192. Branco CM, Rodriguez F. A semi-implicit formulation for compositional reservoir simulation.
SPE Adv Technol Ser. 1996;4:171–7.
193. Arana O. Diffusion and convective mixing of components in naturally fractured reservoirs;
2003.
194. Mansoori J. Discussion of compositional modeling with an equation of state. Soc Pet Eng J.
1982;22:202–3.
195. Watts JW. A compositional formulation of the pressure and saturation equations. SPE Reserv
Eng. 1986;1:243–52.
196. Manzocchi T, Carter JN, Skorstad A, Fjellvoll B, Stephen KD, Howell JA, et al. Sensitivity
of the impact of geological uncertainty on production from faulted and unfaulted shallow-
marine oil reservoirs: objectives and methods. Pet Geosci. 2008;14:3–15. https://doi.org/10.
1144/1354-079307-790.
197. Griewank A, Walther A. Evaluating derivatives : principles and techniques of algorithmic
differentiation, 2nd ed. Society for Industrial and Applied Mathematics; 2008.
276 A. Jamili et al.

198. Matthews JD, Carter JN, Stephen KD, Zimmerman RW, Skorstad A, Manzocchi T, et al.
Assessing the effect of geological uncertainty on recovery estimates in shallow-marine
reservoirs: the application of reservoir engineering to the SAIGUP project. Pet Geosci.
2008;14:35–44. https://doi.org/10.1144/1354-079307-791.
199. Leuangthong O, Deutsch CV. Geostatistics Banff. 2004;1:2004.
200. Warren JE, Skiba FF, Price HS. An evaluation of the significance of permeability measure-
ments. J Pet Technol. 1961;13:739–44. https://doi.org/10.2118/1641-G-PA.
201. Killough JE, Kossack CA. Fifth comparative solution project: evaluation of miscible flood
simulators. SPE Symp. Reserv. Simul., Society of Petroleum Engineers; 1987. https://doi.org/
10.2118/16000-MS.
202. Appleyard JR. Nested factorization. In: Proc. SPE Reserv. Simul. Symp., Society of Petroleum
Engineers; 1983. p. 315–24. https://doi.org/10.2523/12264-MS.
203. Fan L, Harris B, Jamaluddin A. Understanding gas-condensate reservoirs. Oilf Rev 2005;
14–27. https://doi.org/10.4043/25710-MS.
204. Thomas F, Bennion D, Andersen G. Gas condensate reservoir performance. J Can Pet Technol.
2009;48:18–24. https://doi.org/10.2118/09-07-18.
205. Loomis AG, Shea GB. Secondary recovery of oil in California. World Pet. Congr.; 1943.
206. Cook AB, Coulter RH, Spencer GB, Chin T, Elliott WC. Secondary recovery from semide-
pleted oil reservoirs converted to gas-storage operations. Conf. gas supply, Transm. Storage,
Chicago, IL, USA, 10 May 1956; 1956.
Chapter 6
Gas Recycling

Reza Azin, Amin Izadpanahi, and Mohamad Mohamadi-Baghmolaei

Abstract This chapter provides a comprehensive study about gas recycling into the
oil and gas reservoirs. First, the concept of gas recycling is studied. In general, gas
recycling is mainly called to the re-injection of produced gas into the gas condensate
reservoirs in order to maintain the reservoir pressure above the dew point pressure.
But, the re-injection of produced gas into the oil and unconventional reservoirs is also
called gas recycling. This chapter addresses the advantages and disadvantages of this
method, gas recycling mechanisms in fractured reservoirs, the operation design and
well patterns, economical aspects, studies and projects worldwide. Also, a detailed
study of the re-vaporization mechanism during gas injection is investigated. In the
case study, it is shown that gas recycling can significantly improve the gas and
condensate production from a gas condensate reservoir.

Nomenclature

Acronyms

ACS Average Condensate Saturation


AGS Average Gas Saturation
BHP Bottom-hole Pressure (psi)

R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
M. Mohamadi-Baghmolaei
Department of Process Engineering, Memorial University, St. John’s, NL, Canada
e-mail: m.mohammadibaghmolaei@mun.ca

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 277
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_6
278 R. Azin et al.

CCE Constant Composition Expansion


CMG Computer Modeling Group
CRF Condensate Recovery Factor
DECE Designed Exploration and Controlled Evolution
DL Differential Liberation
GE Gas Equivalent
GRF Gas Recovery Factor
QC Quality Check

Variables

A Cross sectional area, ft2


B Formation Volume Factor, reservoir conditions/standard conditions
C Condensate Volume, STB
Cf Formation Compressibility, 1/psi
Cg Bulk gas phase concentration, mol/m3
Cge Equilibrium concentration of the liquid condensate in the gas phase,
mol/m3
Cw Water Compressibility, 1/psi
D Dispersion tensor
Dh Hydrodynamic Dispersion, m2 /s
G Gas Volume, scf
g Gravity acceleration, ft2 /s
H Depth, ft
K0 Lumped mass transfer coefficient
k Permeability, md
krg Gas relative permeability
kro Oil relative permeability
m Ratio of initial gas cap volume to initial oil volume
Mm * Sink/source term representing production/injection of m component,
lbmol/d
Mw Molecular Weight, g/gmol
n Number of Moles
N Oil Volume, STB
P Pressure, psi
q Oil rate, STB/day
R Gas-oil Ratio, scf/STB
Rcnst Ideal Gas Constant, ft3 .psi/lbmol.R
S Saturation
T Temperature, °F
u Interstitial Velocity, m/s
V Darcy Velocity
W Water Volume, bbl
6 Gas Recycling 279

x Mole fraction
xm Molar fraction of the m-th component in the condensate phase
ym Molar fraction of the m-th component in the gas phase

Greek Letters

α Ratio of the volume of produced gas to the volume of injected gas


β Ratio of the volume of recycled gas to the summation of the exited gas
from each separator
φ Porosity
γ Specific Gravity
ρ Density
μi Chemical potential
μ Viscosity

Subscripts

b Bubble
c Condensate
co Component
e Influx
fs First Separator
g Gas
i Initial
inj Injected
l Liquid
o Oil
p Produced
ph Phase
recycled Recycling Operation
s Solution
sc Standard Condition
ss Second Separator
st Stock Tank
t Total
w Water
280 R. Azin et al.

6.1 Introduction

Gas recycling is one of the common methods for improving hydrocarbon recovery
in both oil and gas reservoirs. In this method, all or some part of the produced gas
is re-injected to the reservoir. The main purpose of gas recycling is maintaining
the reservoir pressure. However, it is reported in the literature that gas recycling is
performed in some cases in order to enhance condensate recovery and enhanced oil
recovery. Hydrocarbon gas injection can be considered somehow as the gas recycling
method if the source of injected gas is from the same reservoir.
This chapter provides a comprehensive study of gas recycling. This comprehensive
study addresses the different aspects of the gas recycling operation. These aspects
include the conceptual, economical, historical and operational aspects. Also, a case
study is conducted to show the effects of gas recycling on condensate recovery.

6.2 The Concept of Gas Recycling

Gas recycling is defined as a process in which all or some part of produced hydro-
carbon gases are returned to the reservoirs. This process is common in both gas
condensate and oil reservoirs. The application of gas recycling in oil and gas conden-
sate reservoirs are different. A significant loss of valuable intermediate hydrocarbon
components may remain unproduced as condensate banking in the reservoir as a
result of pressure decline below the dew-point pressure. Hydrocarbon gas injection
keeps the reservoir pressure above the dew point pressure and increases the conden-
sate recovery by revaporizing the condensed liquid in the reservoir [1, 2]. Therefore,
gas recycling has demonstrated to be very effective in decreasing the liquid accumu-
lation in gas condensate reservoir. Also, gas recycling into gas cap of oil reservoirs
keeps the reservoir pressure and minimizes the produced gas oil ratio. Also, it can
produce oil through swelling and decrease residual oil saturation by reducing the
interfacial tension (IFT) [3]. The disadvantage of this method is that the inadequate
mobility ratio can cause gravity override, viscous fingering and decreases the ultimate
hydrocarbon recovery [4].
There has been a debate on distinguishing gas recycling versus gas injection
for pressure maintenance. In some companies, the gas recycling mainly addresses
the gas condensate formations, while gas recycling into oil formation is simply
called “gas injection” or “pressure maintenance”. However, in some projects the
injection of the produced gas into oil reservoirs is called gas recycling as well [5,
6]. Gas recycling expression is also applied to some projects in the unconventional
hydrocarbon reservoirs such as oil shale, condensate shale and heavy oil [7, 8]. More
information about the effect of gas recycling on the recovery of liquids from shales is
available on Fragoso et al. [7]. In heavy oil reservoirs, the aim of using gas recycling
is to enhance oil recovery by decreasing oil viscosity and interfacial tension which
improving oil mobility. For this purpose, natural gas is injected into oil formation and
6 Gas Recycling 281

Fig. 6.1 Schematics of gas recycling project

dissolve in the formation oil under high pressure [8]. Also, re-injection of producing
gas which is the concept of gas recycling is used for miscible gas injection as an
enhanced oil recovery method [9].
Figure 6.1 shows a schematic of a gas recycling project in oil reservoirs. The
produced gas is separated from oil in a single- or multi-stage separator system.
Then, it is pressurized to the desired pressure and reinjected into gas cap. In the case
of partial recycle, part of produced gas flows in the pipeline towards processing unit,
NGL plant or gas refinery.

6.3 Economics of Gas Recycling

The economic is a significant parameter in planning a gas recycling project. The


owner has many options to consume the produced gas, some of which are shown in
Fig. 6.2.
The produced gas may be used for reinjection into gas cap, stored as underground
gas storage, reinjected in an EOR process, used as a feed to petrochemical plant,
or burnt as a fuel in civil and domestic applications like CNG, power plant, water
desalination plant, energy intensive industries like cement, steel, aluminum, and
etc. Alternatively, it can be exported as LNG or PNG. Selection of one or more
applications is based on economic analysis, market demand, and industrial status of
the country.
Figure 6.3 shows the cost elements of a miscible gas injection project reported
by Zekri and Jerbi [10]. Most of the costs which reported in their study are similar
282 R. Azin et al.

Fig. 6.2 Application of associated and natural gas

Fig. 6.3 Costs elements of a miscible gas injection project [10]


6 Gas Recycling 283

to the gas recycling projects. These costs contain fixed and operating costs of sepa-
ration, compression, pipeline transmission, and drilling of injection wells. The main
difference between gas recycling and miscible project is the costs of injection fluid.
In the case of miscible project, CO2 , N2 and hydrocarbon gas can be used based on
the reservoir characteristics and it contains additional costs of capture and transport
of injection. But, produced gas is the main injection fluid in the gas recycling project
which is available on the site of gas recycling project.
However, if the sour gas needs purification before recycling, dehydration and
sweetening units may also be required. In the case of partial recycle, the additional
gas can be used for industrial and domestic applications. In general, the gas recycling
may be applied to balance the gas supply and demand in a region.

6.4 Advantages and Disadvantages of Gas Recycling

Natural gas injection as gas recycling has a number of advantages. Recovery and
storage of produced gas as a natural resource into reservoir and pressure maintenance
of oil reservoir are the main advantages of gas recycling. Control of pressure decline
and water encroachment contributes in stabilized oil production with low produced
gas oil ratio. However, in the case of total recycle, no additional gas will be available
for other applications shown in Fig. 6.2. In other words, the stored gas postpones
gas supply to the market. Also, the costs of installation, operation, and maintenance
of injection facilities should be provided by the company. In addition, the pressure
maintenance potential of recycled gas may be insufficient and the reservoir might
require additional gas to be supplied from external sources. Also, there is a risk of
gas loss through seepage and flow to other formations underground. Table 6.1 shows
the comparison between methane injection and other gases.

Table 6.1 Comparison between methane and other gases [11]


Injected fluid Advantage Disadvantage
Nitrogen -Availability -Formation of early liquid dropout
-Early gas sales -Generation plants costs
-Low compressibility -Separation facilities are required
-High pressure support
Carbon Dioxide -High ability of revaporization - Maintain reservoir pressure partially
-Early gas sales -Not available everywhere
- Reduction of greenhouse gases -Corrosion
-High cost
Methane -High ability of revaporization of -Postpones the gas sales in the market
liquid -High volume of methane is required
due to its high compressibility
-Maintain reservoir pressure partially
284 R. Azin et al.

6.5 History of Gas Recycling

To the best of our knowledge, the first gas recycling project was designed and operated
in 1941 in USA [12]. Since then, many studies have focused on the gas recycling
scheme and its effects on the recovery process in the gas and oil reservoirs. Most of the
researches indicated that gas recycling is a feasible method for pressure maintenance.
In gas condensate reservoirs, gas recycling can maintain the reservoir pressure above
the dew point pressure and prevent the early liquid drop-out in the reservoir. Also,
it can revaporize the condensate bank around the wellbores as it mentioned in the
literature.
As a classic upstream operation, gas recycling has been dealt with in different
levels from laboratory, model simulation, pilot plant operation, to the full-field oper-
ation. In this section, the reported literature about gas recycling is investigated from
the oldest studies to the newest ones. Most of the gas recycling studies are performed
in gas condensate reservoirs. However, there are other gas recycling studies reported
in oil, heavy oil and shale oil reservoirs. Table 6.2 summarizes some of the published
studies on gas recycling. This table specifies the type of reservoir fluid, region of
study, date of the study, study purpose and the corresponding results of each published
studies.

6.6 Worldwide Gas Recycling Projects

There are many gas recycling field projects around the world reported in the liter-
ature. These field projects consist of oil and gas condensate reservoirs. Table 6.3
summarizes the worldwide gas recycling projects and their corresponding results.
Figure 6.4 shows the worldwide gas recycling projects and studies by location.
Also, this figure shows the type of reservoir fluid in each study or field project.

6.7 Reservoir Material Balance Calculations for Gas


Recycling

As discussed, gas recycling projects may apply to oil and gas condensate reservoirs.
The foundations for general material balance equation in both cases will be presented.

6.7.1 Material Balance for Gas Recycling in Oil Reservoir

The general form of volumetric material balance for an idealized container which is
shown in Fig. 6.5 is as follows [43]:
6 Gas Recycling 285

Table 6.2 Studies about gas recycling and corresponding results


Type of Region Published Studied subject Results of the study Ref.
reservoir year
fluid
Gas USA- Louisiana 1946 -Actual well tests Gas recycling [12]
condensate of injected gas approximately
-Dilution curves recovered fifty
for several wells percent of gas
-Analyzing the reserves
core-analyses data
Gas reservoir USA- LA 1966 -The procedure of Managing a high [13]
designing and pressure gas
installing a high recycling without
pressure gas extraordinary costs
recycling system for equipments
Gas USA- Gulf of 1969 -Studied the Development of [14]
condesate Mexico techniques for computer programs
estimating gas to perform
reserves and calculations of
forecasting furure recovery and
performance in a economic evaluation
gas condesate
reservoir under
gas recycling
project
Gas Canada- Alberta 1970 -Evaluating the The ultimate [15]
condensate liquid recovery will be
revaporization by higher with the
lab experiments lower gas sales rate
-Study the cyclic in comparison with
performance by the partial recycling
compositional and no recycling
model simulation
Oil Canada-Alberta 1974 -Compositional Gas recycling is a [16]
simulation of gas suitable recovery
recycling mechanism because
of:
-Vaporize the
condensate in the gas
cap
-Vaporize a large
amount of residual
oil
(continued)
286 R. Azin et al.

Table 6.2 (continued)


Type of Region Published Studied subject Results of the study Ref.
reservoir year
fluid
Gas Not Mentioned 1992 -Simulation the -Gas injection into [17]
condensate effect of gas gas condensate
recycling and N2 causes the rising of
injection on liquid drop-out
condensate -Gas recycling
recovery causes much less
liquid drop-out in
comparison of N2
injection
Gas Algeria- Toual 2003 -Simulation the The benefits of gas [18]
condensate Field recovery of recycling are:
natural depletion -Maintain the
and gas recycling reservoir pressure
-Economical -Prevent liquid
evaluation of gas dropout
recycling project -Higher condensate
recovery in
comparison of
natural depletion
-Increasing the
volume and duration
of gas injection
causes higher
condensate recovery
Gas Algeria- Hassi 2006 -Simulation the Gas recycling: [2]
condensate R’Mel effect of gas -Maintain the
recycling on reservoir pressure
condensate -Increasing the
recovery condensate recovery
by preventing the
liquid dropout
-Injection volume is
an important
parameter in the
amount of recovery
Gas Iran 2006 -Simulation and -In dual permeability [19]
condensate experimental resrvoirs model the
investigation of condensate recovery
the effect of gas due to gas recycling
recycling on in 2.5 times higher
condensate than dual porosity
recovery in reservoirs
fractured
reservoirs
(continued)
6 Gas Recycling 287

Table 6.2 (continued)


Type of Region Published Studied subject Results of the study Ref.
reservoir year
fluid
Gas Iran 2010 -Simulation -CO2 gas injection [20]
condensate comparison of has the higher
natural depletion recovery in
and different comparison of gas
kinds of gas recycling
injection
including gas
recycling
Gas Oman 2014 -Simulation study -Condensate [21]
condensate of gas recycling recovery factor
increases with
increasing gas
recycling ratio
-Condensate
recovery factor is
increased 6% by
40% recycling ratio
for 8 years
Gas Indonesia 2014 - Stochastic - Optimization of [22]
condensate optimization of well placement and
parameters that operational
affecting the gas paramaters can
recycling significantly
improved the net
present vale of
operation
- Condensate
saturation decreased
by optimizaing the
well placement, sales
volume and
production rates
Shale USA- Texas 2015 -Simulation study - Gas recycling [7]
condensate- of gas recycling could improved the
Shale oil and lean gas liquid recovery
injection into
shale condensate
and shale oil
reservoirs
(continued)
288 R. Azin et al.

Table 6.2 (continued)


Type of Region Published Studied subject Results of the study Ref.
reservoir year
fluid
Gas Not Mentioned 2016 -Simulation of -The most important [23]
condensate gas recycling and parameters on
its’ effects on condensate recovery
condensate asre the gas
recovery injection,drawdown
pressure and
operational
parameters
-Fraction of light
components have
negative effect on the
condensate recovery
Oil Nigeria 2017 -Studied five -Re-injecting the [24]
cases for produced gas from
improving oil oil rim to water oil
recovery from contact is selcted as
thin oil rim the best option for
minimizing the
water cut
-This technique
increases the
producing GOR
-The produced gas
should be injected in
the gas cap for
decreasing the GOR
Oil USA 2017 -Simulation study -Recycling a high [25]
of a number of percentage of
scenario in a produced gas,
partially depleted gradual reduction of
reservoir CO2 injection and
addition of new
wells is suggested
for increasing the
final recovery of the
studied field
6 Gas Recycling 289

Table 6.3 Worldwide gas recycling field projects and their corresponding results
Region Level of project Reservoir fluid Results Ref.
Middle East Pilot Oil Reservoir Pressure and [26]
oil production was
improved in the both
studied pilots
Canada Field Gas Condensate Gas cycling is required [27]
because of the
retrograde condensation
behavior of reservoir
fluid
USA- Texas Field Gas Condensate The field results showed [28]
the low recovery
because of early
breakthrough of gas
USA- Gulf of Mexico Field Gas Design and installation [13]
of a gas cycling facility
for high pressure gas
injection
China Field Oil Cyclic gas injection can [8]
improve well production
rate 3–5 times
Indonesia Field Oil Gas recycling is [29]
(Handil EOR Project) identified as a technical
and economical EOR
process and it is applied
into 11 reservoirs
Indonesia Field Gas Condensate Gas recycling was [30]
conducted to sweep the
condensate rich gas
towards the producers
USA- Louisiana Field Gas Condensate Pressure maintenance [31]
achieved by gas cycling
method
Sudan Field Oil Gas recycling delayed [6]
early water
breakthrough
UAE Field Gas Condensate Gas recycling is used to [32]
maintain reservoir
pressure
Middle east Pilot Oil Comprise the results of [3]
secondary and tertiary
gas injection
United Kingdom- North Field Oil Maintaining reservoir [33]
sea pressure with gas
recycling scheme
(continued)
290 R. Azin et al.

Table 6.3 (continued)


Region Level of project Reservoir fluid Results Ref.
Indonesia- Arun Field Gas condensate Gas recycling can delay [1]
the decline of
Condensate-Gas Ratio
(CGR), postpone
increasing Water–Gas
Ratio and prevent liquid
drop out
USA- Texas Field Gas condensate Gas recycling helped [34]
the recovery of heavy
components of reservoir
fluid
Australia-East Timor Field Wet gas [35]
Sea
Canada- Alberta Field Gas condensate The field produced [15]
hydrocarbons just after
the cycling operation
began
South America- Field Oil Gas recycling had the [36]
Colombia best recovery in
comparison of other
development strategy
China Field Gas condensate Pressure maintenance is [37]
achieved by dry gas
re-injection
Canada- Alberta Field Gas condensate Maintaining reservoir [38]
pressure by gas
recycling
USA- Texas Field Oil Gas recycling improved [39, 40]
oil recovery by
vaporizing the immobile
oil in the flow path of
recycled gas
Colombia Field Gas condensate High recovery factor [41, 42]
achieved by the
favorable conditions of
gas recycling which acts
as the miscible
conditions and pressure
maintenance
6 Gas Recycling 291

Fig. 6.4 Worldwide projects and studies of gas recycling (references are given in Tables 6.2 and
6.3)

Fig. 6.5 Idealized container (left) at initial condition, (right) after producing and different
mechanisms activation
292 R. Azin et al.

The parameter m is defining as the ratio of initial volume of gas cap to the volume
of oil initially in place. The definition of the terms of Eq. 6.1 are as follow:
     
N Boi + m N Boi = N − N p Bo + m N Boi /Bgi Bg
      
+ N Rsi − N p R p − N − N p Rs Bg + We − W p Bw
   
+ N Boi (1 + m)((Cw Swi − C f )/(1 − Swi )) p + G r ecycled Bgr ecycled
 
+ Win j Bw (6.1)

Pore volume occupied by the oil initially in N Boi


place
Pore volume occupied by the gas in the gas cap m N Boi
 
Pore volume occupied by the remaining oil N − N p Bo
 
Pore volume occupied by the gas cap at m N Boi /Bgi Bg
reservoir pressure p
   
Pore volume occupied by the evolved solution N Rsi − N p R p − N − N p Rs Bg
gas
Net water influx We − W p Bw
Change in pore volume due to initial water and N Boi (1 + m)((Cw Swi − C f )/(1 − Swi ))p
rock expansion
 
Pore volume occupied by the recycled gas G r ecycled Bgr ecycled
 
Pore volume occupied by the water injection Win j Bw

The original oil in place calculates by rearranging the general material balance
(Eq. 6.1):
   
N p Bo + G p Bg − G recycled Bgrecycled − N p Rs Bg − We − W p Bw − Win j Bw
N=        
(Bo − Boi ) + (Rsi − Rs )Bg + m Boi Bg /Bgi − 1 + Boi (1 + m) Swi cw + c f /(1 − Swi ) p
(6.2)

In this step, a constant will introduces to the Eq. 6.2 which is defined in Eq. 6.3.

G r ecycled Bgr ecycled


α= (6.3)
G p Bg

α shows the ratio of the volume of produced gas to the volume of injected gas.
By introducing this constant to the Eq. 6.2, the material balance equation can be
investigated in the cases of total or partial gas recycling. Equations 6.4 through
6.6 shows the numerator of material balance equation for the reservoir with no gas
recycling, partial gas recycling and total gas recycling, respectively.
   
α=0: N p Bo + G p Bg − N p Rs Bg − We − W p Bw − Win j Bw (6.4)
6 Gas Recycling 293

   
α = cons(0 < α < 1) : N p Bo + G p Bg − αG recycled Bgrecycled − N p Rs Bg − We − W p Bw − Win j Bw
(6.5)
   
α=1: N p Bo − N p Rs Bg − We − W p Bw − Win j Bw (6.6)

The injection term is not available in the case without gas recycling. In the case
of partial injection, both gas injection term and gas production term are available
in the material balance equation. The injection term in this case is multiplied with
the proportion of the volume of injection to the production volume which defined in
Eq. 6.3. In the case of total gas recycling, the produced gas is reinjected into reservoir.
In this case, the injection term equals the gas production terms so, the gas production
term is eliminated as shown in Eq. 6.6.
Based on the characteristic of the reservoir each term could be considered or
could be neglected. For example for a reservoir with no active aquifer and no water
injection the terms We -Wp Bw and Winj Bw are elimiated from the general material
balance equation. Also, the rock and fluid expansion mechanism is neglected due
to little effect on recovery. By eliminating the rock and fluid expansion term and
considering no active aquifer, no water injection and operating partial and total gas
recycling the general material balanced are simplified as follows respectively:

N p Bo + G p Bg − αG r ecycled Bgr ecycled − N p Rs Bg


Partial gas recycling : N=   
(Bo − Boi ) + (Rsi − Rs )Bg + m Boi Bg /Bgi − 1
(6.7)
N p Bo − N p Rs Bg
Total gas recycling : N=   
(Bo − Boi ) + (Rsi − Rs )Bg + m Boi Bg /Bgi − 1
(6.8)

This relation shows the initial oil in place for a reservoir with a solution gas drive
and gas recycling into the reservoir.
294 R. Azin et al.

6.7.2 Material Balance for Gas Recycling in Gas Condensate


Reservoirs

Figure 6.6 shows the gas condensate system in the initial conditions and after

Fig. 6.6 Idealized container


for gas condensate system
(left) at initial condition,
(right) after producing and
bottom water drive activation

producing with the bottom water drive mechanism. The overall material balance
for the mentioned system is expressing as follows:
   
G Bgi = C Bc + (G − G p )Bg + We − W p Bw
   
+ G Bgi ((Cw Swi − C f )/(1 − Swi )) p + G r ecycled Bgr ecycled (6.9)

The terms in Eq. 6.9 are defining as follows:

Pore volume occupied by the gas initially in place G Bgi


Pore volume occupied by the remaining condensate C BC
Pore volume occupied by the gas reservoir pressure p (G − G p )Bg
Net water influx We − W p Bw
Change in pore volume due to initial water and rock G Bgi ((Cw Swi − C f )/(1 − Swi ))p
expansion
 
Pore volume occupied by the recycled gas G r ecycled Bgr ecycled
6 Gas Recycling 295

The revaporized condensates flow with the gas phase towards the production
wells, where they may experience phase change within the well or at the surface. For
high-rate gas wells in lean gas condensate reservoirs, it is shown that the revaporized
condensate has little impact on the flow mechanisms in the wellbore. Also, modelling
of surface facilities like wellhead chokes and wellhead separators are not affected by
the revaporized condensate in lean gas condensate reservoirs [44–46]. The volume
of vaporized condensates may be converted to gas equivalent and added to the gas
stream. Schematic of the two and three stage separation processes are shown in
Fig. 6.7.
The cumulative produced gas is a combination of gas, gas equivalent of conden-
sate, and gas equivalent of water vapor. The total production gas for three and two
stage separation are given by Eq. 6.10 and 6.11, respectively.

G p = G p(sur f ) + G E t = G f s + G ss + G st + G E c + G E w (6.10)

G p = G p(sur f ) + G E t = G f s + G st + G E c + G E w (6.11)

By introducing the equation of three stage separator into Eq. 6.9, the general form
of material balance can be investigated in order of partial or total gas recycling. For
this purpose, a constant is defined as the ratio of the volume of recycled gas to the
summation of the exited gas from each separator. This constant is shown as follows:

G r ecycled Bgr ecycled


β= (6.12)
(G f s Bg f s + G ss Bgss + G st Bst )

Fig. 6.7 Schematic of separation processes (up) three stage separation (down) two stage separation
296 R. Azin et al.

The material balance equation for the case of no gas recycling, partial gas recycling
and total gas recycling are shown in Eqs. 6.13 through 6.15.

G(Bgi − Bg ) = C Bc − (G f s Bg f s + G ss Bgss + G st Bst + C p G E c + W p G E w )


β=0 C Swi −C f
+G r ecycled Bgr ecycled + (We − W p Bw ) + (G Bgi ( w1−S wi
)p)
(6.13)
G(Bgi − Bg ) = C Bc − (1 − β)(G f s Bg f s + G ss Bgss + G st Bst ) + C p G E c + W p G E w
β = cons C Swi −C f
+(We − W p Bw ) + (G Bgi ( w1−S wi
)p)
(6.14)

β =1 G(Bgi − Bg ) = C Bc + C p G E c + W p G E w + (We − W p Bw )
Cw Swi − C f
+ (G Bgi ( )p) (6.15)
1 − Swi

The gas equivalent in the above equations for condensate and water are calculated
by following equations and it is a function of density (or API) and molecular weight:

n Rcnst Tsc γC
G Ec = = 133000 (6.16)
Psc MwC
n Rcnst Tsc
G Ew = = 7390 (SC F/sur f ace bbl) (6.17)
Psc

6.8 Operation Design and Well Pattern in Gas Recycling

The gas recycling projects can be designed and operated in single-layered, multi-
layered, non-fractured, as well as fractured reservoirs. For multi-layered formations,
such parameters as stratification, permeability variation between layers, connectivity
between layers in terms of vertical to horizontal permeability become important.
Also, diffusion of gas from injection layer into other layers can be effective. As an
examples, Sprinkle et al. studied the effect of stratification and variable permeability
on the gas recycling projects. Their studied reservoir is a gas condensate reservoir
which consists of 15 stratifications with the variable permeability in the range of 1.6
to 3836 md. The results of their study show that the gas recycling is not suitable for
the reservoirs with variable permeability and different stratifications [28].
It should be mentioned that in multilayered reservoirs, the horizontal permeability
stratification may result in a premature gas breakthrough. As a consequence, the oil
recovery factor would be low and the reservoir will experience high gas oil ratio at
early time and the whole project is expected to fail [28].
There are two cases to be considered for planning the development of gas recycling
wells. The first case is when the recycling begins after a long period of natural
6 Gas Recycling 297

depletion and the second case the recycling is planned from the discovery of the
field.
In the first case, the reservoir geometry, production history physical characteristics
of reservoir such as the size of gas cap to reservoir oil are known. The gas sweep
efficiency and pressure profile of reservoir can be evaluated by a reservoir model for
existing wells in the field. Then, the maximum sweep efficiency can be determined
for the optimum injection and production rate and possibly by drilling some new
wells.
In the second case, the optimum arrangement of the wells will be selected for the
highest recovery and lowest investment after building a reliable reservoir model and
calculating the well capacities based on the results of production tests. More data for
repeating the model study will be available as the development of the field begins.
There are three guidelines available for the model study as following [47]:
(1) In the case of converging displacement, some place for the injection wells
and particular well rates are considered so as to attain gas breakthrough at the
production wells. In this way, the maximum recovery of heavy components is
achieved.
(2) In the case of long and thin structures, the production wells are placed at one
end of the field and the injection wells at the other. The displacement is linear
in this case and a high sweep efficiency is gained.
(3) In circular reservoirs displacement with unknown aquifer activity, in order to
prevent the production of considerable quantities of water, it is recommendable
to inject the gas towards the center of the structure.
The number of injection wells is an important parameter in the amount of recovery
caused by gas recycling. Balogun et al. studied the gas recycling scheme on a gas
condensate field in Niger-Delta. They compared a base case (natural depletion) with
the three gas recycling scenarios which consist of 1-injector, 2-injector and 3-injector
wells. Their results show that 3-injector is not suitable for gas recycling in the studied
reservoir because of an early gas breakthrough in production wells. There is no
considerable difference between the 2 injector and 1-injector scenario and 1-injector
found to be the optimal scenario for gas recycling in the studied reservoir [48].

6.9 Gas Recycling in Fractured Reservoirs

Recovery mechanisms which improved oil recovery are different in fractured and
unfractured reservoirs. During gas injection in fractured reservoirs, the gas phase
channels through the high permeability fractures. If diffusion is not considered, a
large amount of the oil is flowing through the fractures. The main recovery mech-
anisms in fractured reservoirs are physical diffusion, capillary and viscous forces,
gravity drainage, phase behavior effects and total pore compressibility [49].
298 R. Azin et al.

6.9.1 Gravity Drainage

The density difference between oil and gas is the main cause of gravity drainage.
Gravity drainage has a considerable effect on the oil recovery where the matrix
has low permeability and sufficient height [49]. Gravity drainage may facilitate or
defer the oil recovery during gas injection [50]. Gas-oil gravity drainage is a feasible
recovery mechanism in the case of mixed-wet, naturally fractured reservoirs where
countercurrent imbibition and water injection does not work. Immiscible gravity
drainage can cause high residual oil saturation and low production rates because of the
capillary holdup. In contrast, miscible gravity drainage can improve ultimate recovery
and production rate due to density, viscosity and interfacial tension reduction. Extra
gas recycling is needed for increasing gravity drainage rates where oil and gas mixing
in fractures. However, mixing gas and oil in the matrix improves local and overall
gravity drainage rates because of faster draining oil into the fractures system [51].
One important factor which affects the recovery is the gas injection rate. There is a
critical gas rate which the recovery at this rate is maximum. The competition between
viscous, capillary and gravity forces causes this phenomenon [50]. The maximum
rate of oil production by the gas-oil gravity drainage (GOGD) in the absence of
capillary forces can be calculated as follows [52]:

kkr o
qG OG D = − Aρog g (6.18)
μo

The capillary pressure gradient is added to the potential gradient in the presence of
capillary forces. As observed, the rate of GODG is increased by viscosity reduction
and increasing the difference between gas and oil density. Gas recycling can enhance
the GOGD process in the fractured reservoir due to the high difference between dry
gas and oil density in comparison to CO2 and N2 injection.

6.9.2 Block-To-Block Process

Block to block process and matrix continuity are two important parameters that
related to gravity drainage. Block to block process is defined as re-entering of
drained oil from one matrix into other matrix blocks. Matrix continuity is known
as the incomplete blocks separation which resulting in some flow path continuity
between the blocks [53]. Block to block process plays an important role on devel-
oped miscibility and oil recovery, but its effect on the development of miscibility
and flow behavior is unclear. The oil recovery from the field may significantly be
delayed due to re-infiltration of the recovered oil into another matrix block (block-
to-block process) [54]. Oil pressure decreases due to oil reduction in the matrix,
while capillary pressure increases. A high flow gradient exists between the fracture
and matrix due to the pressure difference at the matrix/fracture surface. So, the oil in
6 Gas Recycling 299

fracture will flow into the matrix. Oil will preferentially remain in the matrix under
the gas gravity drainage process because the oil pressure in the matrix is lower than
oil pressure in the fracture. Therefore, the remaininig oil saturation is high and the
oil recovery is delayed. The degree of oil distribution in fractures and matrix controls
the delay in oil recovery [55].

6.9.3 Capillary Hold-Up

Capillary hold-up zone is known as the area which is unrecoverable because of lower
or equal capillary forces in comparison to gravitational forces. In cases where the
height of capillary hold-up is greater than the block height, the oil recovery from the
block is not likely to achieve [56].
Density and viscosity reduction is occurred when the injected gas become miscible
with the reservoir oil. Ultimate recovery and gas-oil gravity drainage will be enhanced
due to the single phase flow in miscible conditions. Mixing and IFT reduction play
important roles in miscible condition. Oil viscosity reduction, increase in the gas
density in the fractures and oil density reduction in the matrix is caused by the mixing
mechanism. While, increasing in oil relative permeability, no capillary re-imbibition
and no capillary hold-up is caused by IFT reduction [57].

6.9.4 Diffusion and Dispersion

Diffusion is the main oil recovery mechanism in highly fractured reservoirs with
low permeability matrix under a certain temperature and pressure conditions [50].
Molecular diffusion has a small impact on the recovery in the conventional reservoirs
but, in the fractured reservoir it has a significant impact on the recovery due to the
existence of large surface which available for diffusion. Knudsen diffusion, molecular
diffusion and surface diffusion are the three main mechanisms which attributed in
diffusive mass transfer in the porous media. Knudsen diffusion is dominant in the
reservoir with nanometer pore spaces. Surface diffusion describes the interactions
between molecules and surface. Molecular diffusion is created due to concentration
gradients. Diffusion can dominate the viscous flow in the porous media where fluid
velocities are low. Molecular diffusion can improve the oil recovery by gas injection
in fractured reservoirs. Higher liquid recoveries and lower producing GOR can be
achieved by diffusion of gas components from the fracture into the matrix [58].
300 R. Azin et al.

6.10 Alternatives to Gas Recycling

The common method in gas condensate reservoirs for preventing the condensate
formation is the reinjection of produced gas into the reservoir. Siregar et al. showed
that nitrogen injection is a feasible option for gas injection instead of gas recycling.
Injetion of the produced gas is an expensive method and dry gas is not always avail-
able. But, nitrogen is relatively cheap and available everywhere besides, it has suit-
able injection characteristics such as non-corrosive, safe and environment-friendly
[17]. In many parts of the world, gas recycling is less attractive because of the high
price of dry gas. In some cases, the produced gas is mixing with some portion of
nitrogen for achieving the reservoir pressure maintenance and saving some hydro-
carbon produced gas at the same time [59]. CO2 injection can be considered as an
alternative to gas recycling and it can reduce the residual oil saturation. CO2 injection
is a costly method because it is not available everywhere. The reports show that in
the gas condensate fields with high percentage of CO2 , this method can decrease the
liquid-dropout [11].
Figure 6.8 shows the recovery methods of condensate which can select as an
alternative for gas recycling. There are several methods which can increasing the
condensate recovery from gas condensate reservoir. Gas recycling and huff-n-puff
are categorized as the gas injection methods for increasing the condensate recovery.
In this chapter, gas recycling is investigated in details. The concept of huff-n-puff
is available in the Chap. 1. Another method for condensate recovery is the wellbore
treatment by acidizing and hydraulic fracturing [60, 61]. Drilling horizontal wells

Fig. 6.8 Different methods of increasing condensate recovery


6 Gas Recycling 301

can decrease the drawdown pressure which postpones the formation of condensate
around the wellbore [62]. On the other hand, this method improves the condensate
production due to the increase of the surface between well and reservoir [63].
Changing the interfacial tension (IFT) can be considered as one of the effective
methods for increasing the condensate recovery. Solvent treatment and wettability
alteration are two main methods for changing the IFT. Solvent injection reduces the
surface tension between gas and liquid phases which causes the dissolution of conden-
sate in the gas phase. Methanol is one of the suitable and available solvents which uses
to decreasing the IFT [64]. Different chemical compounds are reported for changing
the wettability of reservoir’s rocks in the literature. These compounds consist of
nano-particles and surfactants. The favorable wettability is selected based on the
conditions of the fluid and reservoir for increasing the condensate recovery. More
information about the wettability alteration method are available in Chap. 3.

6.11 Modelling the Gas Recycling into the Gas Condensate


Reservoirs

6.11.1 Governing Equation

The most reliable mathematical method of gas injection modelling especially in


condensate reservoir is compositional approach. In contrary with two previous
models so-called two-component hydrocarbon model, several components are
involved in calculations and modelling. Every existing component in each phase can
be transferred into the contacted neighbour phase. To solve this model, several auxil-
iary equations including valid EOS, capillary and relative permeability correlations
are required. Though, generating a general differential equation for compositional
model is easy, it is strongly complicated to be solved numerically. The compositional
modelling and more details are presented in Chap. 5. The governing compositional
equation is presented in Eq. 6.19.

∂   
n ph n ph n ph
φ xi j ρ j S j = −∇ xi j ρ j V j + ∇φ Di j ∇ρ j xi j i = 1, 2, . . . , n co
∂t j=1 j=1 j=1
(6.19)

The basic assumption of the compositional model is considering equilibrium state


between two phases in pores. This comes from the fact that the required time for
diffusion across the pore length is short enough in such extant that compositions of
two phases remain constant. In this way, the chemical potential concept is employed
as basic auxiliary equation. The equality of chemical potential in two phases is
presented by Eq. 6.20.
302 R. Azin et al.

μi j = μik j = 1, 2, . . . , n ph k = 1, 2, . . . , n ph k = j (6.20)

It is understood from Eq. 6.20 that equilibrium is established when chemical


potential of component i in phase j is equal to the chemical potential of component i
in phase k. In compositional model, in each time step the fluids in contacts are assumed
to be in equilibrium and via a flash calculation the equilibrium composition could be
determined. Depending on block size and time step equilibrium compositions may
vary and consequently accuracy of results is affected. Also, choosing proper EOS
considering fluid composition and their specifications is of great concern. Even, the
existence of water, wax, asphaltene and hydrate is strongly connected to the choosing
finest EOS with the acceptable capability of predicting phase constituents.
By introducing the Darcy equation into Eq. 6.19, the equation for component m
in two phase flow is as follows:
  

kkr o kkrg  
∇ xm ρo (∇ Po − γo ∇ H ) + ym ρg ∇ Pg − γg ∇ H + Mm∗
μo μg
∂   
= φ xm ρo So + ym ρg Sg (6.21)
∂t

6.11.2 Mass Transfer During Re-vaporization of Condensate

As stated earlier, one of the main recovery mechanisms in gas recycling is the revap-
orization of condensate. In this mechanism, the condensate evaporates into the gas
phase and flows to the surface. This mechanism is occurs when the immobile liquid
is in contact with the flowing gas stream. condensate revaporization is the dominant
recovery mechanism of the liquids which trapped in dead-end pore or pore throat.
One of the key parameters in the quality and quantity of condensate recovery is the
mass transfer between gas and liquid phases [65]. The revaporization of condensate
into the gas phase has three stages according to the results of an experimental test
conducted at atomospheric conditions [66]. These stages schematically are shown in
Fig. 6.9. The first stage is the rising stage of liquid concentration in gas phase. The
second stage is the steady state stage. In this stage, the maximum concentration of
liquid in the gas phase is achieved which is the longest stage during the revaporiza-
tion. The concentration of liquid in gas phase is reduced in the last stage. Decreasing
the liquid saturation causes the reduction of liquid concentration in gas phase in
un-steady state stage.
Mass transfer coefficient shows the quantity and quality of the mass transfer
process. Mohamadi-Baghmolaei et al. studied the mass transfer coefficient for the
revaporization during gas injection. They investigated the effects of different gases
(CH4 , CO2 and N2 ), different gas flow rate, different representing fluid (C5 , C6 , C7 )
and different grain size (porosity). Their studies show that methane has the highest
6 Gas Recycling 303

Steady State Stage

Liquid Concentration in Gas Phase

End of Process

Time

Fig. 6.9 Schematic of condensate revaporization stages [67, 68]

mass transfer coefficient with the lightest fluid (C5 ). Also, they modelled the mass
transfer coefficient in three stages. The main equation of their modelling work is a
one dimensional mass balance equation represented in Eq. 6.22.

∂C g ∂ 2Cg ∂C g
φ Sg = φ Sg D h − φ Sg u + K 0 (C ge − C g ) (6.22)
∂t ∂x 2 ∂x

where Ø, Sg , Dh , u, and K 0 stand for the porosity, gas saturation, hydrodynamic


dispersion, interstitial velocity, and lumped mass transfer coefficient, respectively.
In the first stage, they solve Eq. 6.22 in steady state condition by neglecting the
diffusion/dispersion term. Also, in the steady state condition the accumulation term
(concentration variation with time) is zero. In this case, Eq. 6.22 is simplified to
Eq. 6.23 [66].

∂C g K0
u = (C ge − C g ) (6.23)
∂x φ Sg

In the second stage, the diffusion/dispersion and convection mass transfer are
included to estimate the mass transfer coefficient accurately. Their results show that
the diffusion term has a little effect on the magnitudes of the mass transfer coefficient.
In this case, the Eq. 6.22 is simplified to Eq. 6.24. They solved Eq. 6.24 analytically
for calculating the amount of mass transfer coefficient during their experiments [67].

∂ 2Cg ∂C g
φ Sg D h − φ Sg u = K 0 (C g − C ge ) (6.24)
∂x 2 ∂x
In the last stage, they studied the unsteady state mass transfer. The unsteady state
mass transfer occurs during the falling stage that is shown in Fig. 6.9. They solve
Eq. 6.22 analytically and numerically which shows a good match with experimental
304 R. Azin et al.

data. More details of the analytical and numerical solution, experimental procedure
and modelling study are available in Mohamadi-Baghmolaei et al. [68].

6.12 Case Study: Simulation of Gas Recycling

The data for the gas recycling simulation are provided from the third SPE comparative
solution project [69]. These data consist of fluid data and geometry for investigating
the effect of gas recycling on a gas condensate reservoir. The fluid model was built
using the Winprop PVT module of CMG software. Then, the reservoir model was
constructed using the GEM compositional simulator of the same software. In the
following sections, the fluid behavior modelling results and gas recycling results are
investigated.
Other studies are conducted on the data of the third SPE comparative solution
project. As an exmple, Ayala and Ertekin designed an expert system based on these
data which capable of finding the optimum production porotocols for operating a gas
recycling project [70].

6.12.1 PVT Modelling

PVT data in the third SPE comparative solution project belongs to a gas conden-
sate reservoir. These data include composition of reservoir fluid and laboratory
tests, i.e. constant composition expansion (CCE), constant volume depletion (CVD),
and swelling tests. After entering the composition to the software, C7+ is split and
regrouped. The new heavy fraction is C15+ . Table 6.4 shows the composition before
and after the spilliting and regrouping.
The equation of state (EOS) is tuned using the laboratory test data. Peng-Robinson
equation of state is used to predict the fluid behavior. A , B , acentric factor, volume
shift and volume shift coefficient of the groups after FC6 were chosen as the regression
parameters. The results of the are as follows:
CCE. Figure 6.10 shows the model results and experimental data. Experimental
relative oil volume (ROV) and gas compressibility factor are reported in the CCE
data. It is obvious from these figures that the model predicts the experimental data
accurately.
CVD. Retrograde condensate, produced gas, gas and 2 phase compressibility
factor were measured in the constant volume depletion test. The results of matching
the model to the experimental data were shown in Fig. 6.11. The results indicate that
model can predict liquid volume, produced gas and two phase compressibility factor
accurately. But, there are some minor differences between model and experimental
gas compressibility factor.
Swelling Test. This test is performed using the reservoir fluid and prepared lean
gas with the composition reported in Table 6.5.
6 Gas Recycling 305

Table 6.4 Composition of


Before spilliting After spilliting
reservoir Fluid
Component Mole % Component Mole %
CO2 1.21 CO2 1.21
N2 1.94 N2 1.94
CH4 65.99 CH4 65.99
C2 H6 8.69 C2 H6 8.69
C3 H8 5.91 C3 H8 5.91
IC4 2.39 IC4 2.39
NC4 2.78 NC4 2.78
IC5 1.57 IC5 1.57
NC5 1.12 NC5 1.12
FC6 1.81 FC6 1.81
C7+ 6.59 C07 -C09 3.734149
API of the C7+ 51.4 C10 -C12 1.618236
Specific Gravity of the 0.7737 C13 -C14 0.528877
C7+
Molecular Weight of the 140 C15+ 0.708738
C7+

4.7 1.2
1.15
Gas Compressibility Factor

4.2 Model ROV Model Gas Z Factor


Exp. ROV 1.1
Relative Oil Volume

3.7 Exp. Gas Z Factor


Psat 1.05
3.2 1
2.7 0.95

2.2 0.9
0.85
1.7
0.8
1.2 (a) 0.75 (b)
0.7 0.7
800 1800 2800 3800 4800 5800 6800 0 1000 2000 3000 4000 5000 6000 7000
Pressure (psi) Pressure (psi)

Fig. 6.10 Comparison between the model and experimental data in the CCE test a Relative oil
volume b Gas compressibility factor

Figure 6.12 shows the model and experimental swelling factor which are matched
adequately.
P–T Diagram. Figure 6.13 shows the P–T diagram of reservoir fluid. As it shown,
the reservoir fluid is categorized as the gas condensate based on the reservoir condi-
tions in the P–T diagram. The critical point is shown in Fig. 6.13 and cricondentherm
and cricondenbar of this fluid are 484 °F, and 3447 psi, respectively.
306 R. Azin et al.

25
80
Model Liq. Vol.
70 Model Prod. Gas
20 Exp. Liq. Vol.
Exp. Prod. Gas
Liquid Volume

Produced Gas
60
15 50

40
10
30

5 20
(a) 10 (b)
0
0 500 1000 1500 2000 2500 3000 3500 4000 0
0 500 1000 1500 2000 2500 3000 3500 4000
Pressure (psi) Pressure (psi)

2-Phase Compressibility Factor


0.95
0.85
Compressibility Factor

0.9 0.8
Model Gas Z Factor
Exp. Gas Z Factor 0.75
0.85
0.7

0.8 0.65

0.6 Model 2-phase Z


0.75
Exp. 2-phase Z
(c) 0.55
(d)
0.7
0 500 1000 1500 2000 2500 3000 3500 4000 0.5
0 500 1000 1500 2000 2500 3000 3500 4000
Pressure (psi) Pressure (psi)

Fig. 6.11 Comparison between the model and experimental data in the CVD test a) Liquid Volume
b) Produced Gas c) 2 phase compressibility factor

Table 6.5 Composition of


Component Methane Ethane Propane
lean gas
Mole % 94.68 5.27 0.05

3.2

Final S.F.
2.7
Exp. S.F.
Swelling Factor

2.2

1.7

1.2

0.7
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Composition (mole Fraction of injected gas)

Fig. 6.12 Comparison between the model and experimental data in the swelling test
6 Gas Recycling 307

4000
2 Phase Envelope
3500 Critical Point
Reservoir Condition

Pressure (psi) 3000

2500

2000

1500

1000

500

0
0 100 200 300 400 500 600
Temperature (F)

Fig. 6.13 P–T diagram of reservoir fluid

6.12.2 Compositional Numerical Simulation

The general equation of compositional flow was presented in Sect. 6.11. In this
specific example, the governing equation is presented in Eq. 6.21 by Ayala and
Ertekin [71].
This compositional material balance equation will be solved by the simulator
for each component in each grid block of the system. Compositional numerical
simulation is performed using the CMG-GEM. The reservoir model consists of a 4
layer reservoir as shown in Fig. 6.14.
Tables 6.6 and 6.7 show the grid and reservoir properties in the studied reservoir.
The properties of each layer are reported in Table 6.8.
After building the reservoir model, two scenarios were defined to investigate the
effect of gas recycling on production enhancement. One scenario is designed to
produce with gas recycling and the other will operate without gas recycling. Both

Fig. 6.14 Schematic of the studied reservoir


308 R. Azin et al.

Table 6.6 Grid properties of the studied reservoir


Grid Property Value
nX 9
nY 9
nZ 4
DX (ft) 293.3
DY (ft) 293.3

Table 6.7 Reservoir properties of the studied reservoir


Reservoir property Value
Datum depth (ft) 7500
Porosity 0.13
Gas water contact (ft) 7500
Initial pressure at contact (psi) 3550
Water density (lbm/ft3 ) 63
Water compressibility (1/psi) 3 × 10–6
PV compressibility (1/psi) 4 × 10–6

Table 6.8 Properties of each layer


Layer Horizontal Vertical permeability Thickness (ft) Depth to center (ft)
permeability (md) (md)
1 130 13 30 7330
2 40 4 30 7360
3 20 2 50 7400
4 150 15 50 7450

scenarios were designed a 14-year production. In the gas recycling scenario, the
reservoir produces 10 years with gas recycling and 4 years without gas recycling.
Figure 6.15 shows the gas rate, condensate rate and cumulative condensate produc-
tion in scenarios with and without gas recycling. In the gas recycling scenario, the
gas flow rate is constant until the end of run, while the gas flow rate and condensate
production stop after 10 years without gas recycling. As observed, gas recycling
improved condensate production significantly.
Figure 6.16 shows the reservoir pressure for both scenarios. In the case without
gas recycling, the reservoir pressure dropped rapidly until the production stopped.
But, the case with gas recycling showed a lower pressure drop during gas recycling.
Also, in this case, the pressure drop became sharp when gas recycling stops.
Figure 6.17 compares average gas saturation (AGS) and average condensate satu-
ration (ACS) in both scenarios. The average gas saturation is higher in the gas
6 Gas Recycling 309

7.0E+6 1000
900 Without Cycling

Condensate Rate (bbl/day)


6.0E+6
800 With Cycling

5.0E+6 700
Gas Rate (ft3/day)

600
4.0E+6
500
3.0E+6
400

2.0E+6 300
200
1.0E+6 Gas Rate- Without Cycling
Gas Rate- With Cycling (a) 100 (b)
0.0E+0 0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05 2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date Date
Cumulative Condensate Production

3.0E+6
Without Cycling
With Cycling
2.5E+6

2.0E+6
(bbl)

1.5E+6

1.0E+6

5.0E+5
(c)
0.0E+0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.15 a gas rate b condensate rate c cumulative condensate production in the studied scenarios

4000
Without Cycling
3500
With Cycling
Reservoir Presure (psi)

3000

2500

2000

1500

1000

500

0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.16 Reservoir pressure in the studied scenarios

recycling scenario. The lower condensate saturation in the case with gas recycling
indicates condensate revaporization in the presence of injected dry gas.
Figure 6.18 compares gas recovery factor (GRF) and condensate recovery factor
(CRF) in the scenarios. This figure indicates that the gas recovery factor is much
higher in the scenario without gas recycling, while the condensate recovery factor is
higher in the gas recycling scenario. The gas recovery factor remains low in the gas
recycling scenario until the gas injection stops and then increases.
310 R. Azin et al.

1
0.9
0.8
Average Saturation 0.7
0.6
0.5
AGS- Without Cycling
0.4 AGS- With Cycling
ACS- Without Cycling
0.3 ACS- With Cycling

0.2
0.1
0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.17 Comparison of the gas and condensate average saturation in the defined scenarios

GRF- Without Cycling


100 GRF- With Cycling
CRF- Without Cycling
CRF- With Cycling
80
Recovery Factor (%)

60

40

20

0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.18 Gas and condensate recovery factor

Figure 6.19 shows condensate saturation in two blocks of production well (7-
3-3 and 7-3-4) and a block far from the production well (1-1-3) in the production
with gas recycling and without gas recycling scenarios. According to this figure,
the condensate saturation in the well blocks increases rapidly from the beginning
of production. Also, the effect of gas recycling on condensate saturation is more
highlighted near the wellbore. As observed, condensate saturation in the blocks of
production well decreases by gas recycling. However, changes in the condensate
saturation are small in blocks far from the production well due to low pressure drop.
6 Gas Recycling 311

0.4
Grid 7-3-3- Without Cycling
0.35 Grid 7-3-4- Without Cycling
Grid 1-1-3- Without Cycling
Condensate Saturation Grid 7-3-3- With Cycling
0.3
Grid 7-3-4- With Cycling
Grid 1-1-3- With Cycling
0.25

0.2

0.15

0.1

0.05

0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.19 Comparison of gas saturation for two blocks of production well and a random block far
from the production well in two production scenarios

6.12.3 Optimization

Optimization is conducted using the CMG-CMOST on the case with gas recy-
cling to investigate the effect of production and injection rate, location of injec-
tion well and status of injection well on condensate recovery. This module contains
powerful engines for sensitivity analysis, history matching, optimization and uncer-
tainty assessment. This study uses the CMG DECE engine. This engine is an iterative
optimization method which consists of two different satges, namely, designed explo-
ration stages and controlled evolution stage. Figure 6.20 shows the flow chart of this
optimization method. The DECE method shows a good result for optimization and
assisted history matching in the literature [72–76].
In the optimization process, the location of the injection well and the rate of
injection and production are optimized to obtain higher condensate recovery from
the reservoir. Table 6.9 shows the optimization parameters and the optimized value
of each parameter.
Figure 6.21 shows the QC plot which demonstrates the quality of created proxy.
Proxy here means the approximation which is built by the engine. It is usually a
correlation between the input parameters and the output parameter. The results close
to 45° show the agreement between proxy results and actual simulation results. In a
perfect match, all points should fall on the 45-degree line. The R2 of this plot is 0.95
which indicate the proxy worked accurately.
Figure 6.22 is a tornado chart which shows the effect of each parameter on the
condesate recovery. The positive parameters on the right side have a positive impact
on the condensate recovery and the others has a negative impact on the condensate
recovery. As shown in this figure, the condensate recovery decreases due to the
increase in the grid number in i direction of injection well. In contrast, by increasing
312 R. Azin et al.

Fig. 6.20 Flow chart of optimization using DECE engine

Table 6.9 Optimization parameters and the value before and after optimization
Parameters Grid of injection Grid of injection Injectoion rate Production rate
well in I direction well in J direction
InjILocation InjJLocation InjRate ProdRate
Default value 1 9 4,700,000 6,200,000
Optimized value 8 5 5,170,000 4,650,000
6 Gas Recycling 313

1.3E+8
Experiments
1.2E+8 45 Degree Line
Recovery Factor (%)
1.1E+8

1.0E+8

9.0E+7

8.0E+7

7.0E+7
7.0E+7 8.0E+7 9.0E+7 1.0E+8 1.1E+8 1.2E+8 1.3E+8
Date

Fig. 6.21 QC plot of the built proxy

Maximum 1.29E+08

Minimum 7.26E+07

InjILocation(1, 9) -2.57E+07
Parameter Name

ProdRate(4650000, 7750000) 1.97E+07

InjJLocation(1, 9) 1.96E+07

InjILocation*InjILocation -1.30E+07

InjRate*InjRate 1.00E+07

ProdRate*ProdRate -9.85E+06

InjRate(3525000, 5875000) 9.34E+06

InjJLocation*InjJLocation -8.42E+06

-5.0E+7 0.0E+0 5.0E+7 1.0E+8 1.5E+8


Condensate Recovery ($)

Fig. 6.22 Tornado chart

the grid number in j direction, the condensate recovery is increased. These facts show
that by closing the injection well, the condensate recovery will be decreased.
Figure 6.23 demonstrates the Sobol analysis for determining the sensitivity of a
parameter. The higher the percentage is, the more sensitive will the parameter be.
According to this figure, grid number in the I direction of injection well has the
highest sensitivity, followed by grid number in the J direction of injection well and
production and injection rates.
Figure 6.24 shows the produced condensate rate in the optimization process
314 R. Azin et al.

InjILocation 43.7201
Parameter

InjJLocation 24.8747

ProdRate 24.5757

InjRate 6.8012

0 10 20 30 40 50
Effects (%)

Fig. 6.23 Sobol analysis

1200

1000
Condensate Rate (bbl/day)

Optimal Solution

800 Base Case

600

400

200

0
2019/08 2022/05 2025/02 2027/11 2030/08 2033/05
Date

Fig. 6.24 Produced condensate after optimization

including the base case and the optimal solution of the problem. As observed, the
produced condensate rate was higher in the optimized case compared to the base
case.

References

1. Suhendro S. Review of 20 years hydrocarbon gas cycling in the arun gas field. In: SPE/IATMI
Asia pacific oil gas conference exhibition. Society of Petroleum Engineers; 2017.
6 Gas Recycling 315

2. Adel H, Tiab D, Zhu T. Effect of gas recycling on the enhancement of condensate recovery,
case study: Hassi R’Mel South field, Algeria. In: International oil conference and exhibition in
Mexico. Society of Petroleum Engineers; 2006.
3. Agrawal P, Kumar J, Draoui E. Lesson learnt from immiscible gas injection pilot in offshore
carbonate reservoir. In: Abu Dhabi international petroleum exhibition conference. Society of
Petroleum Engineers; 2016.
4. Carpenter C. Miscible and immiscible gas-injection pilots in a middle east offshore environ-
ment. J Pet Technol. 2016;68:75–7.
5. Cavanagh A, Ringrose P. Improving oil recovery and enabling CCS: a comparison of offshore
gas-recycling in Europe to CCUS in North America. Energy Procedia. 2014;63:7677–84.
6. Tang X, Wang R, Zhang H. Innovative field-scale application of injecting condensate gas and
recycling gas into medium oil pool: a case study in Sudan. In: SPE Enhanced oil recovery
conference. Society of Petroleum Engineers; 2013.
7. Fragoso A, Wang Y, Jing G, Aguilera R. Improving recovery of liquids from shales through gas
recycling and dry gas injection. In: SPE latin American and caribbean petroleum engineering
conference. Society of Petroleum Engineers; 2015.
8. Jing T, Xiao L, Zhao J. Field study of enhancing oil recovery by gas recycling injection in
ultra deep heavy oil reservoirs. In: SPE asia pacific oil gas conference exhibition. Society of
Petroleum Engineers; 2008.
9. Vissers JM, Kindy A, Hassan A. The integration of miscible gas injection in oil reservoirs with
the development of a sour gas condensate field. In: SPE EOR conference oil gas west Asia;
2012:9. https://doi.org/10.2118/154992-MS.
10. Zekri A, Jerbi KK. Economic evaluation of enhanced oil recovery. Oil Gas Sci Technol.
2002;57:259–67.
11. Syzdykov M. Evaluating gas injection performance in very low permeability, thick carbonate
gas-condensate reservoirs to improve ultimate liquid yield. Colorado School of Mines. Arthur
Lakes Library; 2014.
12. Miller MG, Lents MR. Performance of bodcaw reservoir, cotton valley field cycling project;
new methods of predicting gas-condensate reservoir performance under cycling operations
compared to field data. In: Drilling and production practice. American Petroleum Institute;
1946.
13. Stallings LW. Design and installation of a high-pressure gas cycling system. J Pet Technol.
1966;18:1286–90. https://doi.org/10.2118/1381-PA.
14. Butterfield OR, Clark JD, Brauer EB. Reservoir analysis and recovery prediction for cycling
gas-condensate reservoirs. In: Fall meeting of the society of petroleum engineers of AIME.
Society of Petroleum Engineers; 1965.
15. Abel W, Jackson RF, Wattenbarger RA. Simulation of a partial pressure maintenance gas cycling
project with a compositional model, carson creek field, Alberta. J Pet Technol. 1970;22:38–46.
https://doi.org/10.2118/2580-PA.
16. Thompson FR, Thachuk AR. Compositional simulation of a gas-cycling project, bonnie glen
D-3A Pool, Alberta, Canada. J Pet Technol. 1974;26(1):281–285, 294.
17. Siregar S, Hagoort J, Ronde H. Nitrogen injection vs. gas cycling in rich retrograde condensate-
gas reservoirs. In: International meeting petroleum engineering. Society of Petroleum Engi-
neers; 1992.
18. Belaifa E, Tiab D, Dehane A, Jokhio S. Effect of gas recycling on the enhancement of condensate
recovery in Toual field Algeria, a case study. In: Production and operations symposium. Society
of Petroleum Engineers; 2003.
19. Shadizadeh SR, Rashtchian D, Moradi S. Simulation of experimental gas-recycling experi-
ments in fractured gas/condensate reservoirs. In: SPE gas technology symposium. Society of
Petroleum Engineers; 2006.
20. Moradi B, Tangsiri Fard J, Rasaei MR, Momeni A, Bagheri MB. Effect of gas recycling on
the enhancement of condensate recovery in an Iranian fractured gas/condensate reservoir. In:
Trinidad tobago energy resource conference. Society of Petroleum Engineers; 2010. https://
doi.org/10.2118/132840-MS.
316 R. Azin et al.

21. Cobanoglu M, Khayrutdinov F, Linthorst S, Iqbal M. Improving condensate recovery of a rich


sour gas condensate field by gas recycling. In: SPE EOR conference oil gas west Asia. Society
of Petroleum Engineers; 2014.
22. Siddiqui MAQ, Al-nuaim S, Khan RA. Stochastic optimization of gas cycling in gas condensate
reservoirs; 2014:10–3.
23. Temizel C, Kirmaci H, Tiwari A, Balaji K, Suhag A, Ranjith R, et al. An investigation of
gas recycling in fractured gas-condensate reservoirs. In: Abu Dhabi international petroleum
exhibition and conference. Society of Petroleum Engineers; 2016.
24. Ogolo NA, Molokwu VC, Onyekonwu MO. Proposed technique for improved oil recovery
from thin oil rim reservoirs with strong aquifers and large gas caps. In: SPE Niger, annual
international conference and exhibition. Society of Petroleum Engineers; 2017.
25. Ampomah W, Balch R, Grigg RB, Cather M, Gragg E, Will RA, et al. Performance assessment
of CO2 -enhanced oil recovery and storage in the Morrow reservoir. Geomech Geophys Geo-
Energy Geo-Resour. 2017;3:245–63.
26. Kumar J, Agrawal P, Draoui E. A case study on miscible and immiscible gas injection pilots
in middle east carbonate reservoir in an offshore environment. In: International petroleum
technology conference; 2015.
27. Field MB, Wytrychowski IM, Patterson JK. A numerical simulation of kaybob south gas cycling
projects. J Pet Technol. 1971; 23(1):251–253, 262.
28. Sprinkle TL, Merrick RJ, Caudle BH. Adverse influence of stratification on a gas cycling
project. J Pet Technol. 1971;23:191–4. https://doi.org/10.2118/2642-PA.
29. Gunawan S, Caie D. Handil Field: three years of lean gas injection into waterflooded reservoirs.
In: SPE Asia pacific improved oil recovery conference. Society of Petroleum Engineers; 1999.
30. Afidick D, Kaczorowski NJ, Bette S. Production performance of a retrograde gas reservoir: a
case study of the Arun field. In: SPE Asia pacific oil gas conference, vol. SPE 28749. Society
of Petroleum Engineers; 1994, pp. 73–80. https://doi.org/10.2118/28749-MS.
31. Owens H Jr, Kirk E Jr, Brinkley TW. Handling high pressure cycling projects at northwest
branc, Louisiana. J Pet Technol. 1959;11:31–4.
32. Torrens R, Mohamed ME, Al Bairaq A, Kumar A. Integrated asset modeling of a gas condensate
field operating under gas recycling mode. In: Abu Dhabi international petroleum exhibition
conference. Society of Petroleum Engineers; 2014.
33. Miller JH. Production enhancement from topsides changes at an offshore gas cycling project.
In: SPE Annu. Tech. Conf. Exhib., Society of Petroleum Engineers; 1992.
34. Justice WH. Review of Cycling Operations in the La Gloria Field. J Pet Technol. 1950;2:281–6.
35. Ledlow LB, Gilbert WW, Omsberg NP, Mencer GJ, Jamieson DP. Revised Big-Bore Well
Design Recovers Original Bayu Undan Production Targets. SPE annual technical conference
and exhibition. Society of Petroleum Engineers; 2008.
36. Ocampo A, Restrepo A, Cifuentes H, Hester J, Orozco N, Gil C, et al. successful foam EOR
pilot in a mature volatile oil reservoir under miscible gas injection. In: IPTC 2013 international
petroleum technology conference; 2013.
37. Jiao Y, Li B, Li Y, Xie W. The dynamic analysis of dry-gas reinjection in a waterdrive gas-
condensate reservoir-case study: yaha gas-condensate reservoir, Tarim Basin, West China. In:
International oil gas conference exhibition. China, Society of Petroleum Engineers; 2010.
38. Wichert E. The windfall field cycling project. J Can Pet Technol. 1963; 2:6–8.
39. McGraw JH, Lohec RE. The Pickton field-review of a successful gas injection project. J Pet
Technol. 1964;16:399–405.
40. Cook AB, Spencer GB, Bayazeed AF. The role of vaporization in high percentage oil recovery
by pressure maintenance. J Pet Technol. 1967;19:245–50. https://doi.org/10.2118/1646-PA.
41. Hoier L, Cheng N, Whitson CH. Miscible gas injection in undersaturated gas-oil systems. In:
SPE annual technical conference and exhibition; 2004:15. https://doi.org/10.2118/90379-MS.
42. Franco CA, Zabala Romero RD, Zapata Arango JF, Mora E, Botero OF, Candela CH, et al.
Inhibited gas stimulation to mitigate condensate banking and maximize recovery in cupiagua
field. In: SPE international symposium exhibition form damage control; 2012:17. https://doi.
org/10.2118/151575-MS.
6 Gas Recycling 317

43. Ahmed T, McKinney P. Advanced reservoir engineering. Elsevier; 2011.


44. Lak A, Azin R, Osfouri S, Gerami S, Chahshoori R. Choke modeling and flow splitting in a
gas-condensate offshore platform. J Nat Gas Sci Eng. 2014;21:1163–70.
45. Azin R, Chahshoori R, Osfouri S, Lak A, HEIDARI SMH, Gerami S. Integrated analysis of
choke performance and flow behaviour in high-rate, deviated gas-condensate wells; 2014.
46. Heidari Sureshjani MH, Azin R, Lak A, Osfouri S, Chahshoori R, Sadeghi F, et al. Production
data analysis in a gas-condensate field: methodology, challenges and uncertainties. Iran J Chem
Chem Eng. 2016;35:113–27.
47. Latil M. Enhanced oil recovery. Éditions Technip; 1980.
48. Balogun O, Alli A, Ogbuli A, Apampa Y, Bahry A, Fasasi R. Evaluating gas recycling scheme in
a retrograde gas condensate field under contractual supply obligations: niger-delta case study.
In: Niger, annual international conference and exhibition. Society of Petroleum Engineers;
2017.
49. Hoteit H, Firoozabadi A. Numerical modeling of diffusion in fractured media for gas injection
and recycling schemes. In: SPE annual international conference and exhibition. Society of
Petroleum Engineers; 2006.
50. Riazi M, Alizadeh AH, Haghighi M, Sample R. Experimental study of gravity drainage during
gas injection in carbonate rocks. Middle East; 2006:1–7.
51. Verlaan M, Boerrigter PM. Miscible gas-oil gravity drainage. In: International oil conference
and exhibition in Mexico. Society of Petroleum Engineers; 2006.
52. Farajzadeh R, Wassing BM, Boerrigter PM. Foam assisted gas–oil gravity drainage in naturally-
fractured reservoirs. J Pet Sci Eng. 2012;94:112–22.
53. Fung LSK. Simulation of block-to-block processes in naturally fractured reservoirs. SPE Reserv
Eng. 1991;6:477–84.
54. Uleberg K, Høier L. Miscible gas injection in fractured reservoirs. In: SPE/DOE improved
oil recovery symposium. Society of Petroleum Engineers; 2002. https://doi.org/10.2118/751
36-MS.
55. Sani RR, Afsari M, Amani M. Effects of block to block interaction on oil recovery from south
pars oil layer. In: SPE project facilities challenges conference METS. Society of Petroleum
Engineers; 2011.
56. van Golf-Racht TD. Fundamentals of fractured reservoir engineering. vol. 12. Elsevier; 1982.
57. Chordia M, Trivedi JJ. Diffusion in naturally fractured reservoirs-a-review. In: SPE Asia pacific
oil gas conference exhibition. Society of Petroleum Engineers; 2010.
58. Shojaei H, Jessen K. Diffusion and matrix-fracture interactions during gas injection in fractured
reservoirs. In: IOR 2015–18th European symposium on improved oil recovery; 2015.
59. Gohary M El, Bairaq AM Al, Bradley DC, Saeed Y. Comparison of condensate recovery by
hydrocarbon and nonhydrocarbon injection. In: Abu Dhabi international petroleum exhibition
conference. Society of Petroleum Engineers; 2015.
60. Suleimenova A, Wang X, Zhu D, Hill AD. Comparative study of acid fracturing and propped
hydraulic fracturing for a tight carbonate formation. In: SPE European features 78th EAGE
conference exhibition. Society of Petroleum Engineers; 2016.
61. Settari A, Bachman RC, Hovem KA, Paulsen SG. Productivity of fractured gas condensate
wells-a case study of the Smorbukk field. SPE Reserv Eng. 1996;11:236–44.
62. Dehane A, Tiab D, Osisanya SO. Comparison of the performance of vertical and horizontal
wells in gas-condensate reservoirs. In: SPE annual technical conference and exhibition. Society
of Petroleum Engineers; 2000.
63. Graf T. Vertical and horizontal Integration to overcome extreme operational challenges for
the achimov tight, gas-condensate formation. In: SPE Russian oil and gas exploration and
production technical conference and exhibition. Society of Petroleum Engineers; 2014.
64. Du L, Walker JG, Pope GA, Sharma MM, Wang P. Use of solvents to improve the productivity of
gas condensate wells. In: SPE annual technical conference and exhibition. Society of Petroleum
Engineers; 2000.
65. Alzate GA, Franco CA, Restrepo A, Del Pino Castrillon JJ, Barreto Alvares DL, Escobar
Murillo AA. Evaluation of alcohol-based treatments for condensate banking removal. In: SPE
318 R. Azin et al.

international symposium exhibition, form: damage control. Society of Petroleum Engineers;


2006.
66. Mohamadi-Baghmolaei M, Azin R, Osfouri S, Zendehboudi S. Evaluation of mass transfer
coefficient for gas condensates in porous systems: experimental and modeling. Fuel. 2019;
255:115507.
67. Mohamadi-Baghmolaei M, Azin R, Osfouri S, Zendehboudi S. Experimental and modeling
investigation of non-equilibrium condensate vaporization in porous systems: effective deter-
mination of mass transfer coefficient. Fuel. 2020; 262:116011.
68. Mohamadi-Baghmolaei M, Azin R, Osfouri S, Zendehboudi S, Hajizadeh A, Izadpanahi A.
Mass transfer during transient condensate vaporization: experimental and modeling study. J
Mol Liq 2020:114022.
69. Kenyon D. Third SPE comparative solution project: gas cycling of retrograde condensate
reservoirs. J Pet Technol. 1987;39:981–97.
70. Ertekin T. Neuro-simulation analysis of pressure maintenance operations in gas condensate
reservoirs. J Pet Sci Eng. 2007;58:207–26.
71. Ayala LF, Ertekin T. Analysis of gas-cycling performance in gas/condensate reservoirs using
neuro-simulation. In: SPE annual technical conference and exhibition. Society of Petroleum
Engineers; 2005.
72. Mirzabozorg A, Nghiem L, Chen Z, Yang C, Hajizadeh Y. History matching saturation and
temperature fronts with adjustments of petro-physical properties; SAGD case study. In: SPE
Kuwait international petroleum exhibition conference. Kuwait; 2012, pp. 10–2.
73. Nguyen NTB, Chen Z, Dang CTQ, Nghiem LX, Yang C, Bourgoult G, et al. Integrated modeling
for assisted history matching and robust optimisation in mature reservoirs. In: SPE/IATMI Asia
pacific oil gas conference exhibition. Society of Petroleum Engineers; 2015.
74. Dang C, Nghiem L, Nguyen N, Yang C, Chen Z, Bae W. Modeling and optimization of
alkaline-surfactant-polymer flooding and hybrid enhanced oil recovery processes. J Pet Sci
Eng. 2018;169:578–601.
75. Al-Mudhafar WJ, Rao DN, Srinivasan S. Robust optimization of cyclic co2 flooding through
the gas-assisted gravity drainage process under geological uncertainties. J Pet Sci Eng.
2018;166:490–509.
76. Izadpanahi A, Azin R, Osfouri S, Malakooti R. Optimization of two simultaneous water and gas
injection scenarios in a high GOR Iranian Oil Field. 82nd EAGE annual conference exhibition,
vol. 2020. European Association of Geoscientists & Engineers; 2020, pp. 1–5.
Chapter 7
Design of Subsurface and Surface
Facilities for Gas Injection

Reza Azin and Ahmad Banafi

Abstract Proper design of surface facilities in a gas injection project is of great


importance from both engineering and economic points of views. A process engi-
neer frequently deals with standards, protocols and codes in designing facilities
available in the literature. In this chapter, engineering facets and basic designs of gas
injection surface facilities including pipeline, compressor, intercooler and separator
were introduced and discussed. A step by step procedure for initial design of different
facilities were shown, and the fundamental concepts and equations for design of each
facility were presented, supporting with case studies in each section so as to ensure
the design effectiveness. All of the case studies were extracted from the field data,
and thus, predicting these data with reliable agreement proves the proper design of
facilities.

Nomenclature

Q Gas flow-rate, standard ft3 /day [SCFD] or m3 /day


f Dimensionless friction factor
P Pressure, Psia or KPa
G Gas gravity (air = 1.0)
T Temperature, R or K
L Pipe segment length, miles or Km
Z Gas compressibility factor at average gas temperature, dimensionless
D Pipe inside diameter, in or mm
E Pipeline efficiency

R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Banafi
Iranian Gas Transmission Company, District 10, Bushehr, Iran
e-mail: ahmad.banafi@postgrad.curtin.edu.my

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 319
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_7
320 R. Azin and A. Banafi

H Elevation, ft or m
μ Gas viscosity, lb/ft.s and poise
ε Absolute pipe roughness, in
γ Adiabatic or isentropic exponent, ratio of specific heats of gas, Cp /Cv
Re Reynolds number
Cp Specific heats of gas at constant pressure
Cv Specific heats of gas at constant volume
Wa Adiabatic work, ft.lb/lf of gas or J/Kg of gas
HP Compressor horsepower
M Gas mass flow-rate, lb/min
H Compressor head, ft.lb/lb
η Compressor efficiency, %
Z Compressibility of gas at suction condition, dimensionless
ηa Compressor adiabatic (isentropic) efficiency
ṁ g Gas mass flow-rate, lbm /hr or Kg/hr
ṁ a Air mass flow-rate per fan, lbm/hr or Kg/hr
NFan Number of fans
t Ambient air temperature, °F or K
N Fan speed under simulation conditions, RPM
Nd Fan speed under cooler design conditions, RPM
q Heat transfer rate, BTU/hr or W
U Overall heat transfer coefficient, BTU/hr.ft2 .°F or W/m2 .K
A Heat transfer area, ft2 or m2
LMTD
Log mean temperature difference, °F or K
F Temperature correction factor
h Inside (gas-side) heat transfer coefficient based on tube outside surface area
rf Combined fouling resistance, hr.ft2 .°F /BTU or m2 /K.W
rm Metal resistance, hr.ft2 .°F /BTU or m2 /K.W
N Bay Number of bays under simulation conditions
Ft Temperature correction factor
ut Settling velocity, m/s
ρv Vapor phase density, Kg/m3
ρL Liquid phase density, Kg/m3
DV Minimum vessel diameter, m

7.1 Introduction

When a gas injection project is to be implemented in a given field or reservoir, the


gas needs be filtered, dewatered, compressed, transported to the wellhead and then
injected into well to reach the target depth of reservoir. Design of surface facility and
injection well are crucial steps in gas injection operation. If the source of gas is far
from reservoir, other processes like gas boosting through compressor station(s) and
7 Design of Subsurface and Surface Facilities for Gas Injection 321

transmission pipeline need to be properly designed as well. In addition, when a gas


injection process is far away from source of power, a power generation unit will be
required. Another point which directly affects process design is the source of gas,
i.e. whether it comes from an oil reservoir as associated gas or a dry gas is available
in the area. This issue was discussed in Chap. 1. A process engineer frequently deals
with standards, protocols and codes in designing facilities available in the literature.
This chapter will focus on the engineering aspects of surface facilities.

7.2 Conceptual Design and General Layout

Compared to other processes, the surface facilities used for gas injection are rather
simple and involve pumping of liquid hydrocarbon (in case miscible injection is to
be implemented), pipeline transportation of liquid and gas, and gas compression.
The primary factor in the conceptual design of surface facility is the source of gas,
followed by an injection scheme, i.e. immiscible or miscible gas injection. Depending
on the source, type, and composition of gas, different scenarios are defined. An
associate gas stream flows into a natural gas liquids (NGL) plant, where liquids are
separated from gas stream. Then, dry gas is compressed and transported to injection
site. For a dry gas, wet gas, and gas condensate, a gas sweetening unit may be
necessary to separate acid gas before injection. Then, the gas will be compressed
and transported to the injection point. Figures 7.1 and 7.2 show general layout of
surface facility for associate gas and dry gas injection. If the NGL and gas sweetening

Associate gas
To NGL

Stock
1 1stst step 2nd step
2st Step
Tank
Step

Res. Fluid

Dry Gas Injec on


Compression

Associate
NGL
Gas

Fig. 7.1 General block diagram of associate gas injection


322 R. Azin and A. Banafi

Gas to refinery

Gas 1st 2st


Condensate Seperator Seperator

Slug
Dehydra on
Catcher

Amine Dew Point


Sweatening Se ng

Injec on Compression

Fig. 7.2 General block diagram of dry gas injection

are considered as independent processes, as it is the case for many fields, surface
facilities for the whole injection process become simple.
When a miscible gas injection is applied, dry gas should be mixed with liquid
hydrocarbon before injection. In this case, dry gas is compressed and liquid hydro-
carbon is pumped, both to reach the injection pressure. The liquid hydrocarbon may
be either NGL, condensate, or LPG. The streams are then mixed before injection into
reservoir. The volumetric ratio of liquid and gas streams are selected based on PVT
studies that optimize oil recovery from reservoir. Normally, a safety factor is taken
into account and the liquid/gas volume ratio is overdesigned to ensure miscibility
conditions throughout reservoir. Figure 7.3 shows schematic of miscible injection.
In the case of a miscible gas injection in Iran, dry gas is injected with NGL to make a
miscible stream at 235 bar. The gas stream is supplied from a refinery around 700 km
away from target reservoir, and NGL is supplied from a nearby NGL plant. To make
sure that miscible conditions prevail at all points in reservoir, a higher NGL/dry
gas volumetric ratio is selected for mixing before injection. In cases where there is
insufficient NGL supply available, e.g. due to alternative NGL consumption in petro-
chemical plant to produce chemicals, gas injection stops to prevent gas segregation
in reservoir.
The temperature of inlet stream into compressor raises as it passes through
different stages. Passing the heated stream through intercoolers cools down the gas
7 Design of Subsurface and Surface Facilities for Gas Injection 323

85 bar
AGHAR
32"-10"
235 bar
Gas 6"
To injec on
235 bar

NGL Equivalent to compressor


30 bar outlet pressure

8000 bbl/day

Fig. 7.3 Schematic of miscible injection

and the whole process approaches isothermal conditions. Meanwhile, any formed
liquid can be separated from intercooler outlet stream in knock out drum, as no
liquid should enter the compressor. This is shown schematically in Fig. 7.4.
If a sour gas, flue gas, or acid gas is to be injected into reservoir, excessive corro-
sion is expected to occur in the pipeline, transportation facilities and equipment.
Therefore, corrosion-resistant materials should be selected and anti-hydrate chemi-
cals should also be used to prevent excessive corrosion and reduce hydrate potential
of acid gas components. The highly corrosive sulfur-containing injection gas requires
higher capital investment with relatively high monitoring and inspection costs in the
well and surface facilities. More information about corrosion in different gas injection
scheme provide in Chap. 9.

rd
3 Separa on Gas

A 2nd Separa on Gas


Schock Tank Gas
C1
V2 C2
HE1 V3 C3
HE2 C4 +D
HE3 To NGL Unit
D V4

C
1st
B A
2nd
V: Vessel
3rd
C: Compressor
Tank
HE: Heat exchanger

Fig. 7.4 Schematic of inter-stage cooling and dehydration


324 R. Azin and A. Banafi

7.3 Design Considerations

The process engineer needs a set of information to start designing surface facilities
for gas injection. The basic information includes general, non-technical information
such as climate data, i.e. yearly average rainfall, maximum daily rainfall, average
relative humidity, minimum and maximum relative humidity in summer and winter
conditions, yearly temperature profile as well as maximum and minimum tempera-
ture, extreme maximum and extreme minimum temperatures, sunshine temperature,
design temperature for air cooler design, wind speed and direction, elevation above
the sea level, average and highest barometric pressure, exact location of project, earth-
quake and seismic potential, history and intensity, etc. Environmental data include
wildlife, protected zones, surface and subsurface water resources, marine and aquatic
life, gardening, agriculture and animal husbandry, as well as socioeconomic and
cultural status of neighboring society. Technical information for design of a gas
injection facility includes source and composition of gas and liquid hydrocarbon,
depth, temperature and pressure profile of candidate reservoir, miscibility conditions
(for miscible gas injection), etc. The source and quality of energy is another factor
that should be considered (Table 7.1 [1]).
The design and operation of surface facilities for gas injection seem similar for
all applications. However, a closer look reveals features which make gas injection
processes different. For example, the UGS operation is associated with sharp varia-
tions in pressure during I/W cycles. In a small UGS reservoir in Iran, the minimum
pressure at the end of withdrawal reaches 2500 psig, while the maximum pressure
at the end of injection period exceeds 5300 psig. Such a pressure difference in a
relatively short period must be handled by surface facility properly designed and
customized against pressure stress. On the other side, the AGI requires materials of
construction that are corrosion resistant.

7.4 Design Basis

The basis for surface facility process design includes physicochemical and transport
properties, phase behavior, composition, and thermal properties of injection fluid and
reservoir oil. These data are partly obtained from technical data books, handbooks,
and partly from laboratory measurements. In addition, capacity of surface facility, i.e.
average, minimum, and maximum gas flow rates as well as composition of streams are
considered as a basis for material and energy balance calculations and detailed design
of surface facility equipment. In addition, units of all calculations and measurements
(SI, cgs, British, etc.) are specified as a basis for all calculations. Moreover, codes
and standards of engineering design should be pointed in the design basis. The
capacity and composition of waste streams should be known as the basis for design of
waste treatment facility. Mechanical design requires set points as basis, e.g. corrosion
allowance, average ambient temperature for insulation design and minimum ambient
7 Design of Subsurface and Surface Facilities for Gas Injection 325

Table 7.1 Design criteria


General Process Mechanical Civil/Structural
Design life Economic analysis Design parameters- Design codes and
Process condition- procedures
Design loads
Meteorological and Piping criteria: Design codes and Project layout
other local data Pressure drop, procedures
Line sizing,
Pipe routing,
Design pressure
Environmental Draining & venting Piping system design Access
requirements philosophy
Operating and Silica deposition Pipes General civil
maintenance criteria construction
Cost minimization Insulation Valves Thermal ponds
Control valve types Fittings Retaining walls
Pressure relief devices Vessels Foundation design
Pumps Mechanical Structural design
equipment loads
Other components Pipe supports &
anchors
System isolation Constructability and Structures
philosophy maintainability
Instrument air- source Concrete design
and materials
Sampling & testing Steel design
requirements

temperature for non-insulated equipment. Another temperature set point, i.e. design
basis is required for personnel protection and safety measures. More information
on design basis for different heat transfer, mass transfer, and fluid flow processes
and equipment design may be found in [2–4]. For economic evaluation and analysis
of the whole process, economic information such as the lifetime of surface facility,
crude oil and injection gas prices, cost of energy, electricity, and steam, interest and
inflation rates, etc. should be set as economic basis. These information are used
for profitability analysis and economic evaluation of surface facility such as rate of
return, net present value, etc. The next sections deal with basic design of gas injection
facilities, i.e. pipeline, compressor, intercooler, and separator.
326 R. Azin and A. Banafi

7.5 Pipeline Design

Design of piping network and branches depend on the composition of injection


gas, operating pressure, and injection flow rate. A comprehensive pipeline design
involves hydraulic design, fittings, mechanical design, and economic evaluation. The
pipeline used for transmission of injection gas to wellhead is designed according to
engineering procedures and standards, e.g. ASME B31.8 [5]. It should be noted
that the design criteria and materials of construction is different for transmission
and injection of hydrocarbon and non-hydrocarbon gases. Also, HSE regulations are
stricter for sour gas and AGI applications.
The basic energy equation of flow in pipes is written as follows Eq. (7.1):
 
dP d u2 g
− = + dZ + dF (7.1)
ρ 2gc gc

The energy loss due to friction can be calculated by Eq. (7.2):

f L u2
F= (7.2)
D 2gc

The differential form of friction loss is given by Eq. (7.3):

f u2d L
dF = (7.3)
2gc D

For gases, changes in static and velocity head are small and can be neglected.
Therefore, Eq. (7.1) simplifies to Eq. (7.4) which takes into account friction loss
only.

f u2d L
d P = −ρ (7.4)
2gc D

Substituting the definition of density for gases Eq. (7.5) in Eq. (7.4) gives pressure
profile in the gas pipeline.

ρ = z RT /P M (7.5)

In Eq. (7.5), compressibility factor of gas can be determined through equation


of state (EOS), generalized correlations or by intelligent techniques described in
Chap. 2. Mohamadi Baghmolaei et al. [6] proposed a virial-type EOS for modelling
pipeline and compressor design of gas boosting system, which is quite similar to gas
injection surface facility.
As seen from the above mentioned equations, for a given pipe diameter and length,
it is possible to predict gas flow rate as a function of pressure along the pipeline.
7 Design of Subsurface and Surface Facilities for Gas Injection 327

There are several equations available for calculating gas flow-rate based on gas
properties, pipe diameter and length as well as downstream and upstream pressures
of a pipe segment. The basic equation for this purpose is general flow equation which
is developed for steady-state isothermal gas flow in pipeline as follows [7]:
  0.5
Tb P12 − P22
Q = 77.54 D 2.5 (U SC S units). (7.6)
Pb GT f L Z f
  0.5
Tb P12 − P22
Q = 1.1494 × 10−3 D 2.5 (S I units). (7.7)
Pb GT f L Z f

where:
Q: gas flow-rate, standard ft3 /day [SCFD] or m3 /day.
f: dimensionless friction factor.
Pb : base pressure, Psia or KPa.
Tb : base temperature, R or K.
P1 : upstream pressure, Psia or KPa.
P2 : downstream pressure, Psia or KPa.
G: gas gravity (air = 1.0).
Tf : average gas temperature, R or K.
L: pipe segment length, miles or Km.
Z: gas compressibility factor at average gas temperature, dimensionless.
D: pipe inside diameter, in or mm.
There is also some equations that consider a constant friction factor. It is transpar-
ently obvious that this assumption could not be a realistic one. At least it is expected
that friction factor be a function of pipe diameter. Take Weymouth equation as an
example of such equation which is applicable for gas pipelines with large diameter
and high pressure and gas flow-rate. This equation is formulated as follows [7]:
  0.5
Tb P12 − es P22
Q = 433.5E D 2.667 (U SC S units) (7.8)
Pb GT f L e Z
  0.5
−3 Tb P12 − es P22
Q = 3.7435 × 10 E D 2.667 (U SC S units) (7.9)
Pb GT f L e Z

where
E: pipeline efficiency.
Le : equivalent length of pipe segment, miles or Km

L(es − 1)
Le = (7.10)
S
328 R. Azin and A. Banafi

 
H2 − H1
S = 0.0375G (U SC S units) (7.11)
Tf Z
 
H2 − H1
S = 0.0684G (S I units) (7.12)
Tf Z

where
S: dimensionless elevation adjustment parameter.
H1 : upstream elevation, ft or m.
H2 : downstream elevation, ft or m.
The equivalent friction factor for Weymouth equation is:

4
f = 2 (U SC S units) (7.13)
11.18D 1/6
4
f = 2 (S I units) (7.14)
6.521D 1/6

Reynolds number is defined by Eq. (7.15):

ρu D uD
Re = = (7.15)
μ ϑ

It is difficult to calculate velocity and density of a fluid in industry context, espe-


cially for gases. For a compressible fluid, the velocity can be estimated from flow
rate, area, and the density obtained from EOS. By substituting these parameters in
Eq. (7.15) and with some manipulations, the Reynolds number can be calculated
using the following expression [7]:
  
Pb GQ
Re = 0.0004778 (U SC S units) (7.16)
Tb μD
  
Pb GQ
Re = 0.5134 (S I units) (7.17)
Tb μD

where
μ: gas viscosity, lb/ft.s and poise.
Other symbols and dimensions are as defined earlier.
It is a better approach to calculate Reynolds number with these equations, because
it is essentially related to the flow-rate, a parameter which is easier and more
accurately measured.
The friction factor, f, is a dimensionless parameter and can be estimated based
on Reynolds number. Friction factor can be obtained by Moody diagram, shown in
7 Design of Subsurface and Surface Facilities for Gas Injection 329

Fig. 7.5 Moody diagram

Fig. 7.5 or empirical correlations. Many empirical correlations have been proposed
so far in order to calculate the friction factor. The Colebrook-White equation is one
of the most popular relations amongst them formulated as follows:
   

1/ f = −2 log10 (ε/3.7D) + 2.51/ Re f f or Re > 4000 (7.18)

where
ε: absolute pipe roughness, in.
Re: Reynolds number.
The pipe roughness shown on the right vertical axis, E/D, is determined from
standard charts for different pipe materials [2].
In addition, the friction loss due to valves, fittings, junctions, are determined by
the following equation:

 u2
F = 2/ f = K (7.19)
2gc

Constant K is available for all fittings in reference books, e.g..


Case Study 7.1: Design of a gas pipeline1 Using the Weymouth equation, calculate
the flow-rate in a gas pipeline, 263.244 km long, with outside diameter of 1422.4 mm
and wall thickness of 22.225 mm at an efficiency of 0.95. Upstream and downstream
pressures are 8580 kPa and 7000 kPa, respectively. Compressibility factor is assumed

1 The data for this case study are extracted from field data of a gas pipeline in Iran.
330 R. Azin and A. Banafi

to be 0.881. The gas gravity, flowing temperature, base pressure and base temperature
are respectively considered 0.62, 307.15 K, 101.56 kPa and 288.7 K.
Neglect elevation differences along pipe. Assume a pipe roughness of 0.0018 in.
Compare the calculated flow-rate with measured flow-rate of 62.5 MMSCMD.

Solution: For the given outside diameter and thickness, the inside diameter can be
calculated:
D = 1422.4 − 2 × 22.225 = 1375.95 mm
The flow rate is determined through Eq. (7.9)
  2 s 2 0.5
P1 −e P 2
Q = 3.7435 × 10−3 E TPbb GT f L e Z
D 2.667 = 3.7435 ×
 288.7  0.5
85802 −70002
10−3 (0.95) 101.56 0.62×307.15×263.244×0.881
(1375.95)2.667
Q = 56.25M M SC M D

7.6 Compressor Design

Compressors are used to increase the gas pressure and compensate pressure loss
during gas transmission while providing the required injection pressure underground.
The gas pressure at the compressor outlet must be high enough to compensate friction
losses through pipelines, valves and junctions as well as well head and tubing losses.
According to Fig. 7.6 a balance between compressor outlet and injection pressure
can be written as follows:

Pout − Ppi pe − Pvalve − PW H − Pwell = Pin j (7.20)

Keeping the compressor fluid at single phase is essential throughout all compres-
sion stages, and any liquid must be removed from the gas stream before entering
any stage of compressor. In addition, compression of a gas raises its temperature

Fig. 7.6 Schematic of Pin


pressure drop during gas
injection
Pout ΔPline ΔPW.H.
ΔPvalve

ΔPwell

Pinj
7 Design of Subsurface and Surface Facilities for Gas Injection 331

V1 C1
V2 C2
HE1 V3 C3
HE2
HE3
V4
V: Vessel C: Compressor HE: Heat exchanger

Fig. 7.7 Schematic of compression unit

and it is necessary to cool down the outlet stream of a compressor stage prior to
entering next stage to prevent damage to compressor shaft, bearing, coupling, etc.
To do this, the outlet stream passes through an intercooler followed by a knock out
drum (K.O. Drum) for liquid removal from gas before entering next compression
stage. Schematic of the whole process is shown in Fig. 7.7.
Generally, centrifugal compressors are used to pressurize gas for injection. In the
case of AGI processes characterized by low flow rate and high discharge pressure,
reciprocal compressors are used. In AGI, the feed pressure is approximately 35–
70 kPa which may raise up to 15 MPa at the injection point [8].
A general schematic of energy balance through a compressor as a control volume
is shown in Fig. 7.8.
Based on Fig. 7.8, the energy balance for compressor is written as follows:

Ẇs = ṁ(h o − h i ) + Q loss (7.21)

The lost heat, Qloss , can be reduced to approach towards an adiabatic process. An
adiabatic process with no friction is named an isentropic process. An ideal compressor
is the one that compresses the gas through an isentropic process which implies
equality of entropy for inlet and outlet stream:

Si = So (7.22)

Fig. 7.8 Schematic of hi


compressor energy balance

ho

Qloss
332 R. Azin and A. Banafi

Fig. 7.9 Adiabatic


compression path

From a thermodynamic point of view, gas pressure and volume in an adiabatic


compression are related as following equation:

P V γ = Const. (7.23)

where
γ: adiabatic or isentropic exponent, ratio of specific heats of gas, Cp /Cv .
Cp : specific heats of gas at constant pressure.
Cv : specific heats of gas at constant volume.
Figure 7.9 illustrates an adiabatic compression path on a thermodynamic diagram.
The polytropic process is another path for gas compression in which the pressure–
volume is similar to adiabatic process. However, unlike the isentropic process, a
compressor in the polytropic process is subject to heat transfer to the surroundings.
In such process, gas pressure and volume can be related as follows:

P V n = Const. (7.24)

where n is polytropic exponent.


In the design of compressors, usually an ideal form, i.e. isentropic (adiabatic with
no friction) or polytropic, is selected as the basis for which the required work is
determined. The required work for compressing 1 lb (or 1 kg) of gas through an
adiabatic path can be calculated through the following equation [9]:
  γ −1
53.28 γ P2 γ
Wa = T1 − 1 (U SC S units) (7.25)
G γ −1 P1
  γ −1
286.76 γ P2 γ
Wa = T1 − 1 (S I units) (7.26)
G γ −1 P1
7 Design of Subsurface and Surface Facilities for Gas Injection 333

where
Wa : adiabatic work, ft.lb/lf of gas or J/Kg of gas.
T1 : suction temperature of gas, R or K.
P1 : suction pressure of gas, psia or KPa.
P2 : discharge pressure of gas, psia or KPa.
Other parameters and dimensions are as mentioned earlier. The compressor power,
then, can be calculated as follows [9]:

M × H
HP = (7.27)
η

where
HP: compressor horsepower.
M: gas mass flow-rate, lb/min.
H: compressor head, ft.lb/lb.
η: compressor efficiency, %.
More common formulation for compressor power that considers the gas compress-
ibility is as follows [9]:
       γ γ−1
γ Z1 + Z2 1 P2
H P = 0.0857 QT1 − 1 (U SC S units)
γ −1 2 ηa P1
(7.28)
  
    γ γ−1
γ Z1 + Z2 1 P2
Power = 4.0639 QT1 − 1 (S I units)
γ −1 2 ηa P1
(7.29)

where
Power: compression power, KW.
Q: gas flow-rate. MMSCFD or million m3 /day.
Z1 : compressibility of gas at suction condition, dimensionless.
Z2 : compressibility of gas at discharge condition, dimensionless.
ηa : compressor adiabatic (isentropic) efficiency.
In addition to compressor adiabatic (isentropic) efficiency, a mechanical efficiency
ηm of the compressor driver should be considered so as to acquire the brake HP (BHP)
or brake power required to run the compressor as follows:

HP
BH P = (7.30)
ηm
334 R. Azin and A. Banafi

Power
Brake Power = (7.31)
ηm

where HP or power are calculated from the Eqs. (7.28) and (7.29).
As mentioned earlier, in the design of a compressor, either an isentropic process or
a polytropic process may be selected as a basis. The ideal power and head calculated
for these two processes will be different. Once the ideal power is divided by its
corresponding efficiency, e.g. polytropic ideal work divided by polytropic efficiency,
the actual power obtained by each path will be the same [10, 11].
The compression ratio is defined as the ratio of outlet and inlet pressure:

Pout
CR = (7.32)
Pin

For multistage compressors with equal CR at all stages, compression ratio per
stage is defined as:
  n1
Pout
C R stage = (7.33)
Pin

As a general rule, CR should be less than or equal 4 for a proper compressor


operation.
Finally, the discharge temperature of the gas through an adiabatic process can be
calculated from the following equation:
   γ γ−1
Z1 P2
T2 = T1 (7.34)
Z2 P1

Thermodynamic principles governing compressor design are found in most


classical textbooks and design materials, e.g. [12].
Case Study 7.2: Power and outlet temperature of a gas compressor2 Calculate the
discharge temperature for an adiabatic compression of 36.67 MMSCMD gas with
inlet temperature of 44.26 °C as well as suction and discharge pressures of 968
psia and 1320 psia, respectively. Assume gas compressibility factor at suction and
discharge condition and the adiabatic exponent to be, respectively, 0.9129, 0.9294
and 1.482.
Compare the calculated discharge temperature with the design discharge temper-
ature of 72.19 °C.
Calculate the required compression power if the adiabatic efficiency is assumed
to be 0.8.
Solutions: From Eq. (7.34), the discharge temperature is:
 0.9129  1320  1.482−1
T2 = (44.26 + 273.15) 0.9294 968
1.482
= 344.87 K = 71.72 ◦ C

2 The data for this case study are taken from a gas compressor station in Iran.
7 Design of Subsurface and Surface Facilities for Gas Injection 335

The required compression is calculated by Eq. (7.29).


 1.482    1  1320  1.482−1

Power = 4.0639 1.482−1 ×36.67×317.41 0.9129+0.9294


2 0.8 968
1.482
− 1 =
17774 KW

7.7 Intercooler Design

Air cooled heat exchanger (ACHE), also known as intercooler in compression


process, is commonly used to reduce temperature of inter-stage outlet gas stream
using ambient air and also to decrease the risk of equipment damage and corrosion
at high temperature. Normally, the upper limit of discharge temperature is suggested
by compressor vendor and is selected as the basis for design of intercooler. In addi-
tion, cooling after compression liquefies part of heavy components in the gas stream
which are separated in the K.O. drum right after intercooler.
In a typical ACHE, the hot process fluid flows through heat conducting tubes
and the ambient air is circulated vertically by fan(s) across the tubes. When the tube
section is located on the discharge side of the fan, the ACHE is known as forced draft.
Oppositely, if the tube section is located on the suction side of the fan, the ACHE
is classed as induced draft. Schematic diagrams of these two types of ACHE are
shown in Figs. 7.10 and 7.11, respectively. To improve the efficiency of heat transfer
in an ACHE, the tubes are provided generally with external fins which significantly
increase the heat transfer surface. This extended surface not only compensates for
low heat transfer rate of ambient air at atmospheric pressure, but also helps to operate
the system at a low enough velocity in order to reduce the fan power consumption.
Assembly of tubes, headers and frames in a typical ACHE is named tube bundle.

Fig. 7.10 Schematic diagram of forced draft ACHE


336 R. Azin and A. Banafi

Fig. 7.11 Schematic diagram of induced draft ACHE

In addition, one or more tube bundles, serviced by two or more fans, including the
structure, plenum and other attendant equipment is called bay in a typical ACHE
[13]. Some typical configurations of fan bays in ACHE are illustrated in Fig. 7.12.

Fig. 7.12 Some typical configurations of fan bays in ACHE


7 Design of Subsurface and Surface Facilities for Gas Injection 337

The procedure for design of air cooled heat exchanger is found in classical heat
transfer text books, e.g. [14] as well as design books, e.g. [14–17].
The general equations to design ACHE are heat and mass balances equations
for both air and the gas [18]. The amount of heat rejected from the hot gas can be
calculated as follows:

qg = ṁ g C pg (T1 − T2 ) (7.35)

where
qg : heat rejected from the hot gas, BTU/hr or W.
ṁ g : gas mass flow-rate, lbm /hr or Kg/hr.
Cpg : specific heat of gas, BTU/lbm . °F or Kj/Kg.K
T1 : hot gas inlet temperature, °F or K.
T2 : hot gas outlet temperature, °F or K.
On the other hand, the amount of heat absorbed by the ambient air is represented
by the following equation:

qa = ṁ a C pa N Fan (t2 − t1 ) (7.36)

where
qa : heat absorbed by ambient air, BTU/hr or W.
ṁ a : air mass flow-rate per fan, lbm/hr or Kg/hr.
Cpg : specific heat of air at ambient pressure and temperature conditions,
BTU/lbm.°F or Kj/Kg.K
NFan : number of fans.
t1 : ambient air inlet temperature, °F or K.
t2 : ambient air outlet temperature, °F or K.
The air mass flow rate per fan, ṁ a , can be calculated as follows [19]:

N t1d
ṁ a = ṁ ad (7.37)
N d t1

where
ṁ ad : fan air mass flow-rate under cooler design conditions, lbm /hr or Kg/hr.
N: fan speed under simulation conditions, RPM.
Nd : fan speed under cooler design conditions, RPM.
t1d : ambient air temperature under cooler design conditions, °R.
t1 : ambient air temperature under simulations conditions, °R.
In addition to Eqs. (7.35) and (7.36), the basic heat transfer equation applied for
all heat exchangers can also be employed for air-cooled heat exchangers as follows:

q = U × A × F × L MT D (7.38)
338 R. Azin and A. Banafi

where
q: heat transfer rate, BTU/hr or W.
U: overall heat transfer coefficient, BTU/hr.ft2 .°F or W/m2 .K
A: heat transfer area, ft2 or m2 .
F: temperature correction factor.
LMTD: log mean temperature difference, °F or K.
In order to calculate the heat transferred from hot gas to ambient air from
Eq. (7.38), the different parameters in this equation must be computed.
LMTD can be computed as following equation:

(T1 − t2 ) − (T2 − t1 )
L MT D = (7.39)
ln TT12 −t2
−t1

Gas Processors Suppliers Association [18] presented two figures in order to obtain
the temperature correction factor, F. One can acquire this factor from Figs. (8–10)
and (9–10) in [18]. To use these diagrams, the temperature-corrected functions P and
R are calculated as follows:
t 2 − t1
P= (7.40)
T1 − t1
T1 − T2
R= (7.41)
t2 − t1

To acquire the overall heat transfer coefficient, U, the following equation may be
employed [19]:

1
U= (7.42)
1
hi
+ r f + rm + 1
ho

where
h i : inside (gas-side) heat transfer coefficient based on tube outside surface area,
BTU/hr.ft2 . °F or W/m2 .K
r f : combined fouling resistance, hr.ft2 .°F /BTU or m2 /K.W
rm : metal resistance, hr.ft2 .°F /BTU or m2 /K.W
h o : outside (air side) heat-transfer coefficient, BTU/hr.ft2 .°F or W/m2 .K.
Typical overall heat transfer coefficients in ACHEs for finned tube surfaces are
provided in Table 7.2.
In Eq. (7.42), the parameter h i can be computed by the following equation [19]:
 0.8
N Bays,d ṁ g
h i = h id . (7.43)
N Bay ṁ gd
7 Design of Subsurface and Surface Facilities for Gas Injection 339

Table 7.2 Overall heat


U [W/m2 .K (BTU/hr.ft2 .°F)]
transfer coefficients in ACHE
for finned tube surfaces [13] Gas cooling service
Air or flue gas at 3.45 bar ~3.4 (~0.6)
(g)
Air or flue gas at 6.9 bar (g) ~6.8 (~1.2)
Air or flue gas at 6.9 bar (g) 9.6–14.1 (1.7–2.5)
Hydrocarbon gases at 5.6–13 (1.0–2.3)
1.034–3.45 bar(g)
Hydrocarbon gases at 11.3–19.8 (2.0–3.5)
3.45–17.23 bar(g)
Hydrocarbon gases at 19.8–30 (3.5–5.3)
17.23–103.4 bar(g)
Liquid cooling service
Fuel oil 6.8–10.2 (1.2–1.8)
Light gas oil 17–23.2 (3.0–4.1)
Light hydrocarbons 22.7–31.7 (4.0–5.6)
Heavy gas oil 14.1–17 (2.5–3.0)
Condensing service
Light hydrocarbons 22.7–31.7 (4.0–5.6)

where
h id : inside heat transfer coefficient under cooler design conditions, BTU/hr.ft2 .°F
or W/m2 .K
N Bays,d : number of bays under cooler design conditions.
N Bay : number of bays under simulation conditions.
ṁ gd : gas mass flow-rate under cooler design conditions, lbm /hr or Kg/hr.
Also, the parameter h o may be obtained as follows [19]:
 
ṁ a 0.6
h o = [h od10 − Ft (t1 − t10 )] (7.44)
ṁ a10

where
h od10 : outside heat transfer coefficient at 10 °C, BTU/hr.ft2 .°F or W/m2 .K
Ft : temperature correction factor.
t1 : ambient air temperature under simulation conditions, °F or K.
t10 : ambient air temperature of 10 °C, °F or K.
ṁ a10 : fan air mass flow-rate at 10 °C, lbm /hr or Kg/hr.
The metal resistance, rm , is [19]:
340 R. Azin and A. Banafi

Table 7.3 Typical values of


rf (hr.ft2 .°F/Btu)
fouling factors, rf ,
(hr.ft2.°F/Btu) [16] Natural gas processing streams
Natural gas 0.001
Overhead vapor products 0.001–0.002
C3 or C4 vapor (condensing) 0.001
Lean oil 0.002
Rich oil 0.001
LNG and LPG 0.001
Process gas streams
Hydrogen 0.001
Acid gases 0.002-0.003
Stable distillation overhead products 0.001
Service gas streams
Carbon dioxide 0.002
Flue gases 0.005–0.01

A0 lnr0 /ri r0 r0
rm = = ln (7.45)
2π K L K ri

where K is thermal conductivity of tube metal (BTU/hr.ft.°F).


The combined fouling resistance, r f , is a constant value. The typical values of
fouling factors are presented in Table 7.3. The data in this table are only representative
of available ones in the open literature. More details and typical values for fouling
factors can be found in heat transfer text books, e.g. [16, 20].
In order to solve a gas cooler problem when the gas outlet temperature (T2 ) is
unknown, the following procedure can be used [19]:
1. Assume a value for T2
2. Calculate the heat rejected from the hot gas (qg ) from Eq. (7.35).
3. Calculate the air mass flow rate per fan (ṁ a ) from Eq. (7.37).
4. Calculate the ambient air outlet temperature (t2 ) from Eq. (7.36) (qa = qg ).
5. Calculate log mean temperature difference (LMTD) from Eq. (7.39).
6. Calculate the temperature-corrected functions (P and R) from Eqs. (7.40) and
(7.41).
7. Determine temperature correction factor (F) from Figs. 7.13 or 7.14.
8. Calculate the inside heat transfer coefficient (h i ), outside heat transfer coeffi-
cient (h o ) and overall heat transfer coefficient (U) from Eqs. (7.43), (7.44) and
(7.42), respectively.
9. Calculate the heat transfer rate (q) from Eq. (7.38)
10. If q calculated in step 9 and qg calculated in step 2 are not equal, the gas outlet
temperature (T2 ) must be changed until q and qg match.
7 Design of Subsurface and Surface Facilities for Gas Injection 341

Fig. 7.13 Conceptual PT


diagram mapped by
successive
compression-cooling paths

Fig. 7.14 Schematic


diagram of a knock-out drum

Case Study 7.3: Heat transfer rate and hot gas outlet temperature in an ACHE3 A
hot natural gas stream with a flow-rate of 1,148,000 kg/hr and temperature of 73.53 °C
is to cool with 41 °C ambient air in a one pass cross-flow forced draft ACHE. The
ACHE consists of 4 bays and 2 fans per bay. The air mass flow-rate per fan is
645105 kg/hr. Specific heat of air and natural gas are, respectively, 1.006 kJ/Kg.°C
and 2.661 kJ/Kg.°C. Assuming overall heat transfer coefficient and heat transfer

3 The design data for this case study are taken from a gas compressor station in Iran.
342 R. Azin and A. Banafi

area to be, respectively, 24.82 W/m2 .°C and 39,441 m2 , calculate the gas outlet
temperature and heat transfer rate.
Compare the calculated gas outlet temperature and heat transfer rate with the
design values of 55 °C and 15,696 KW, respectively.

Solutions: In order to solve this problem, based on the procedure mentioned at the
end of this section, a trial and error method must be employed.

1. Assume hot gas outlet temperature T2 = 60 °C


2. Calculate heat removed from the hot gas (qg ) form Eq. (7.35)

qg = 1148000 × 2.661 × (73.53 − 60) = 41331822 Kj/hr

3. Calculate the ambient air outlet temperature (t2 ) from Eq. (7.36) (qa = qg )

N Fan = number o f bays × number o f f ans per bay = 4 × 2 = 8

41331822
t2 = 41 + = 48.96 ◦ C
645105 × 1.006 × 8

4. Calculate log mean temperature difference (LMTD) from Eq. (7.39)

(73.53 − 48.96) − (60 − 41)


L MT D = = 21.66 ◦ C
ln 73.53−48.96
60−41

5. Calculate the temperature-corrected functions (P and R) from Eqs. (7.40) and


(7.41)

48.96 − 41
P= = 0.245
73.53 − 41
73.53 − 60
R= = 1.7
48.96 − 41

6. Determine temperature correction factor (F) from Figs. 7.13 or 7.14

Since the ACHE in this case is one pass-cross flow, the Fig. 7.13 is used. With
calculated P and R and from Fig. 7.13:

F = 0.97

7. Calculate the heat transfer rate (q) from Eq. (7.38)

q = 24.82 × 39441 × 0.97 × 21.66 = 74061031


7 Design of Subsurface and Surface Facilities for Gas Injection 343

8. Since the calculated q in Sect. 7.7 and calculated qg in Sect. 7.2 are not equal,
the assumption for hot gas outlet temperature must be changed and the above
procedure be repeated until q and qg match.

After repeating the above mentioned procedure and using trial and error method,
with the hot gas outlet temperature of 55.12 ◦ C the q and qg match. Therefore the
solution for this case study would be: T2 = 55.12 °C.
Heat transfer rate = 15,624.8 KW.

7.8 Separator Design

A conceptual 2-phase diagram mapped on the successive compression-cooling paths


shown in Fig. 7.13 indicates the points where the process falls in the two-phase
region and liquid phase is expected to form. In the case of AGI, the liquid phase
may be water at early stages of compression and liquefied acid gas at later stages
[8]. However, in the case of hydrocarbon gas injection, the liquid phase may be
either water or in less proportion, intermediate hydrocarbon. In any case, the liquid
formed during successive compression and cooling must be removed before the gas
enters the next stage. Usually a K.O. drum is used to remove liquid droplets and mist
from gas stream. Schematic of a K.O. drum is shown in Fig. 7.14. Proper design
of a gas–liquid separator depends on the type, flow rate, and physical properties of
fluids to be separated, as well as operating pressure and temperature. The separation
of liquid from gas is, in most cases, the same as solid particles’ separation and
thus, the same techniques and equipment are applied [21]. The two-phase separators
may be designed either horizontal or vertical. However, when the gas–liquid ratio is
high, such as the case in a compression-cooling procedure in a gas injection project,
vertical separators are preferred [18, 22]. Separation vessels generally contain four
sections which are shown schematically in Fig. 7.14. Svercek and Monnery proposed
step-by-step procedures for process and mechanical design of two-phase separators
[22].
Where small traces of liquid in stream can be tolerated, gravity settling in hori-
zontal and vertical vessels would be adequate. In order to increase the efficiency of
separating vessels, where the high efficiencies are required or the droplets are likely
to be small, down to 1 μm, knitted mesh demisting pads (knitted wire mesh mist
extractor) are often used. These pads are available in a wide range of materials, thick-
ness and densities. In the case of liquid separators, stainless steel pads with thickness
and nominal density of 100 mm and 150 kg/m3 , respectively, are frequently used.
Smaller vessels could be used if a demister pad is considered in the vessel design.
Cyclone separators are also frequently employed for gas–liquid separation purposes.
The design procedure for these cyclones is analogous to the gas solid cyclones. To
avoid picking up of liquid from the cyclone surfaces, the inlet velocity should be held
under 30 m/s [21]. In the case of separators in a gas compression station in Iran, the
gas is routed through suction scrubbers located upstream of the compressors to guard
344 R. Azin and A. Banafi

against the inadvertent existence of liquid. These Suction scrubbers are double stage
separators; the first stage is a multi cyclo-tube type separator, which comprises 535
numbers of 2” cyclo-tubes and the second stage is a coalescer type filter separator,
which comprises of 51 numbers of removable polypropylene coalescer filters.
Considering forces acting on liquid droplets in a vertical Knock-out drum, the
liquid droplets will settle out when the gravity force is higher than drag force. In
design of separating vessels, the following equation can be employed in order to
estimate the settling velocity [21]:

ρ L − ρv
ut = K (7.46)
ρv

where
u t : settling velocity, m/s.
ρv : vapor phase density, Kg/m3 .
ρ L : liquid phase density, Kg/m3 .
K = 0.07 if there is a demister pad, and K = 0.15 × 0.07 if there is no demister
pad.
In case of a vertical separator, the vessel diameter must be determined large enough
to decrease the gas phase velocity to an amount less than the settling velocity at
which the liquid droplets will settle out. Therefore, the minimum allowable diameter
is calculated as follows [21]:

4Q v
DV = (7.47)
π ut

where
DV : minimum vessel diameter, m
Q v : gas phase flow-rate, m3 /s.
Figure 7.15 provides preliminary requirements for the vessel height. The height
of the vessel above the gas inlet should be sufficient enough to allow for separation
of liquid droplets. A height equal to vessel diameter or 1 m, whichever is greater,
would be suitable. In addition, for the height between liquid level at the bottom of
vessel and gas inlet, an amount equal to half of vessel diameter or 0.6 m, whichever is
greater, should be used. Furthermore, the liquid level depends on the liquid hold-up
time which is essential so as to have a smooth operation [21]. For separations in
compressor suction or interstage scrubbers typically 10 min for liquid hold-up time
would be allowed [21, 22]. It should be noted that if the required liquid depth is small,
it should be increased in design stage in order to have enough space for installing
the level controller. Finally, if there is any demister pad in separator, a height of at
least 0.4 m should be added to the vessel height. As can be seen from Fig. 7.15, the
vessel height (HV ) is calculated as follows:
7 Design of Subsurface and Surface Facilities for Gas Injection 345

Fig. 7.15. Preliminary


requirements for the vessel
height in a vertical separator

HV = h D + h g1 + h g2 + h L (7.48)

The parameters in right side of the equation are shown in Fig. 7.15.
More details about design of gas–liquid separators can also be found in
engineering books in the field of oil and gas industry, e.g. [18, 23].
Case Study 7.4: Preliminary design of a compressor suction separator4 A Syngas
compressor suction separator with demister pad is used to separate liquid loads in
order not to enter the compressor. Make a preliminary design and determine the size
of this vertical separator to handle 390,500 kg/hr of gas and 66,780 kg/hr of water
at temperature and pressure of 69.3 °C and 21.8 barg. Assume gas and water density
to be 9.054 kg/m3 and 973.9 kg/m3 , respectively.
Compare the calculated vessel diameter and height with the design values of 4.2 m
and 8.3 m, respectively.
Solution: From
 Eq. (7.46), the settling velocity is:
u t = 0.07 973.9−9.054
9.054
= 0.72 m/s
Gas flow-rate, Q v = 390500
9.054×3600
= 11.98 m3 /s

4 The design data for this case study are taken from a petrochemical plant in Iran.
346 R. Azin and A. Banafi

Then, the
 minimum allowable diameter from Eq. (7.47) is:
DV = 4×11.98
π×0.72
= 4.6 m

66780
Liquid f low − rate = = 0.019 m3 /s
973.9 × 3600

Considering liquid hold-up time to be 10 min, the liquid volume held in the vessel
is calculated as:

Liquid volume held in vessel = 0.019 × 10 × 60 = 11.43 m3

Therefore, the liquid depth in the vessel is:

liquid volume held in the vessel 11.43


hL = = = 0.69 m
vessel cr oss sectional ar e π × DV2 /4

Based on Fig. 7.15 and the explanation mentioned in this section, the minimum
vessel height is estimated from Eq. (7.48) as follows:

HV = h D + h g1 + h g2 + h L = 0.4 + 4.6 + 2.3 + 0.69 = 7.99 m

7.9 Well Design

Proper design, completion, monitoring, and management of a gas injection well


is crucial for sustainable gas injection project and requires careful consideration
of operating conditions and application. Like all applications, well design for gas
injection involves casing design, cement design, drilling mud design, completion,
and materials of construction.
The operating conditions of a gas injection may differ from one case to another.
Also, formulation of cement and selection of casing material require composition
and operating conditions of injecting gas which may differ in each case. For example,
gas injection in a cyclic operation like UGS poses cyclic stress on casing and cement
behind it, which increases the risk of gas loss through wellbore and behind the
casing. Shahvali et al. addressed defects in the cement that can causes the loss of
bond between cement-casing and cement-formation, and increase the risk of gas
transmission through the space behind casing. These defects include casing and
formation lack in surface roughness, cement volume shrinkage, thin mud layer or mud
channels presence at the contact surface, free water channel or layering in the deviated
wells, and excessive thermal, hydraulic, and mechanical stresses at the wellbore [24].
Hiagh addresses well design differences between depletion and CO2 injection gas
wells [25]. He mentioned different commercial alloys used for well completion in
7 Design of Subsurface and Surface Facilities for Gas Injection 347

CO2 injection projects, including super Duplex (25% Cr with tungsten, GRE lined
carbon steel, Duplex, carbon steel, and S725CrW. Benge suggested SM-2550, high
nickel, and molybdenum as proper alloys resistant for AGI [26, 27]. On the other
side, injection of acid gases, CO2 , and sour gas increases the risk of corrosion in well.
Improper well design may result in a complete failure which can disintegrate well and
make delay in injection/withdrawal [28]. King and King reviewed different causes to
well constructions failure which results in oil and gas leak to the environment. They
reported six failures in gas injection wells. According to their survey, overall leak
frequency ranges between 0.005–0.03% of wells in service [29]. It is well-known that
H2 S is highly corrosive and causes sulfide stress crack propagation in steel casing
[30]. For AGI, Benge [26, 27] addressed challenges in completion liner design for
AGI, the main factor being excessive corrosion in casing, high injection rate, and the
potential for salt zone loading. Records of well failure gas leakage from gas injection
projects like Wabamun Area, Canada [31] may associate with environmental risks
and gas emission as well as loss of inventory [29]. Therefore, the general criteria
of well design are presented here and the reader is suggested to refer the standard
resources for a thorough and detailed discussion for the project under study.
The procedure for formulation of cement is available in classic textbooks based
on reservoir fluid composition, geology, pressure and temperature profile [32]. In
the case of gas injection projects, composition of injecting fluid affects the routine
procedure. For example, CO2 converts calcium silicate in Portland cement to calcium
carbonate which increases cement permeability to and makes it soluble in acid. Benge
suggested the motivation for a proper cement design as little interaction with CO2
and formulating a high-alumina cement system with low fluid loss as CO2 resistant
[26, 27]. In another study, Nygaard et al. indicated low mechanical cement strength as
the cause to increase the risk of failure in cement during CO2 sequestration project in
Wabamun area, Canada [31]. They proposed a workflow to identify wells candidate
for workover to improve leakage integrity and prevent CO2 leakage. In another study,
Fakhreldin et al. proposed a cement formulation without iron to prevent formation
and precipitation of iron sulfide oxide [33].

References

1. Miranda JLH, López LAA. Piping design: the fundamentals. Short Course Geotherm Drilling,
Resour Dev Power Plant Organ by UNU-GTP LaGeo, St Tecla, El Salvador; 2011.
2. Coker AK. Ludwig’s applied process design for chemical and petrochemical plants. Gulf
professional publishing; 2014.
3. Towler G, Sinnott R. Chemical engineering design: principles, practice and economics of plant
and process design. Elsevier; 2012.
4. Peters MS, Timmerhaus KD, West RE. Plant design and economics for chemical engineers,
vol. 4. McGraw-Hill New York; 1968.
5. Institute ANS. Gas transmission and distribution piping systems. American Society of
Mechanical Engineers; 2018.
348 R. Azin and A. Banafi

6. MohamadiBaghmolaei M, Mahmoudy M, Jafari D, MohamadiBaghmolaei R, Tabkhi F.


Assessing and optimization of pipeline system performance using intelligent systems. J Nat
Gas Sci Eng. 2014;18:64–76.
7. Menon ES. Transmission pipeline calculations and simulations manual. Gulf Professional
Publishing; 2014.
8. Carroll JJ. Acid gas injection and carbon dioxide sequestration, vol. 42. Wiley; 2010.
9. Menon ES. Gas pipeline hydraulics. CRC Press; 2005.
10. Moshfeghian DM. How to estimate compressor efficiency. John M Campbell Co; 2015.
11. Moshfeghian M. Compressor calculations: Rigorous using equation of state versus shortcut
method, 2011.
12. Campbell JM, Maddox RN. Gas conditioning and processing. vol. 121. Campbell Petroleum
Series; 1974.
13. Standard T, Ministry I. Engineering standard for process design of dryers original edition 2007;
2000.
14. Incropera FP, Dewitt DP, Bergman TL, Lavine AS. Fundamentals of heat and mass transfer,
6th edn (trans: Xinshi G, Hong Y); 2007.
15. Kern DQ. Process heat transfer. Tata McGraw-Hill Education; 1997.
16. Serth RW, Lestina T. Process heat transfer: principles, applications and rules of thumb.
Academic Press; 2014.
17. von Karman T. The analogy between fluid friction and heat transfer. Trans Am Soc Mech Eng.
1939;61:705–10.
18. Mohitpour M, Golshan H, Murray MA. Pipeline design & construction: a practical approach.
American Society of Mechanical Engineers; 2000.
19. Association. GP, (U.S.) GPSA. Engineering data book : FPS version. Tulsa, Okla. (6526 E.
60th St., Tulsa 74145): Gas Processors Suppliers Association; 2004.
20. Cao E. Heat transfer in process engineering. New York: McGraw-Hill; 2010.
21. Coulson JM, Richardson JF, Backhurst JR, Harker JH. Coulson & Richardson’s chemical
engineering. Butterworth-Heinemann; 1996.
22. Svrcek WY, Monnery WD. Design two-phase separators within the right limits. Chem Eng
Prog. 1993;89:53–60.
23. Bahadori A. Natural gas processing: technology and engineering design. Gulf Professional
Publishing; 2014.
24. Shahvali A, Azin R, Zamani A. Cement design for underground gas storage well completion.
J Nat Gas Sci Eng. 2014;18:149–54.
25. Haigh M. Well design differentiators for CO2 sequestration in depleted reservoirs. Offshore
Eur.: Society of Petroleum Engineers; 2009.
26. Benge G, Dew EG. Meeting the challenges in completion liner design and execution for two
high rate acid gas injection wells. SPE Drill Complet. 2006;21:180–4.
27. Benge G. Cement designs for high-rate acid gas injection wells. In: International petroleum
technology conference; 2005.
28. Singh BK. Well completion challenges in an extreme environment: a case study of the design,
material and completion equipment selection process for extreme environments. In: Abu Dhabi
International petroleum technology conference, society of petroleum engineers; 2015.
29. King GE, King DE. Environmental risk arising from well-construction failure–differences
between barrier and well failure, and estimates of failure frequency across common well types,
locations, and well age. SPE Prod Oper. 2013;28:323–44.
30. Grimes WD, French RN, Miglin BP, Gonzalez MA, Chambers BD. The physical chemistry
nature of hydrogen sulfide gas as it affects sulfide stress crack propagation in steel. Corros.
2014, NACE International; 2014.
31. Nygaard R, Salehi S, Weideman B, Lavoie RG. Effect of dynamic loading on wellbore leakage
for the wabamun area CO2 -sequestration project. J Can Pet Technol. 2014;53:69–82.
7 Design of Subsurface and Surface Facilities for Gas Injection 349

32. Mitchell RF, Miska SZ, Aadnoy BS, Adams N, Barker JW, Cunha JC, et al. Directional drilling.
Fundam Drill Eng Ed Mitchell RL Miska, SZ, SPE Textb Ser; 2011, 12.
33. Fakhr Eldin Y, Irvine-Fortescue J, Grieve J, Taoutaou S, Jain B, Al Kalbani S, et al. The use
of specialized cement to ensure long term zonal isolation for sour wells in South Oman. In:
International petroleum technology conference; 2009.
Chapter 8
Water-Hydrocarbons System

Amin Izadpanahi and Reza Azin

Abstract In this chapter, several methods will be discussed which can be used to
predict the water content of gases. Most of these methods are based on equations
of state and rigorous thermodynamic models. Different methods of predicting water
content of acid gas systems are evaluated based on the literature experimental data. In
addition, the water content diagrams compatible with the experimental data for pure
CO2 , H2 S, CH4 and their mixture are presented. These charts use for facility type
calculations and trouble shooting. In another section, the gas solubility concepts
will be reviewed. This section contains hydrocarbon gas solubility in water, non-
hydrocarbon gas solubility in water, the calculation of gas solubility in water using
Henry’s law constant and the effect of salinity on gas solubility. At the end of this
chapter, a brief review has been conducted on the activity models.

Nomenclature

Variables

f Fugacity
P Pressure,
R Universal gas constant, cm3 MPa/(kmole)
T Temperature,
Vi Partial molar volume, cm3 /gmole
W Water Content, kg/106 Sm3

A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir
R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 351
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_8
352 A. Izadpanahi and R. Azin

y Mole fraction in gas phase

Greek

μ Chemical potential
γ Activity coefficient

8.1 Introduction

In order to plan a successful gas injection project, some aspect about the phase
behavior of water-hydrocarbon system must take into accounts such as water content
of hydrocarbon gases and the solubility of hydrocarbon and non-hydrocarbon gases
in the water. In this chapter, several methods will be discussed which can be used to
predict the water content of gases. Most of these methods are based on equations of
state and rigorous thermodynamic models. The phase behavior is complicated and
extra care should be taken to assure a correct prediction. Although not addressed
in this section, hydrates can also form and these can significantly complicate phase
behavior. Different methods of predicting water content of acid gas systems are
evaluated based on the literature experimental data. In addition, the water content
diagrams compatible with the experimental data for pure CO2 , H2 S, CH4 and their
mixture are presented. These charts use for facility type calculations and trouble
shooting. In another section, the gas solubility concepts will be reviewed. This section
contains hydrocarbon gas solubility in water, non-hydrocarbon gas solubility in water,
the calculation of gas solubility in water using Henry’s law constant and the effect of
salinity on gas solubility. At the end of this chapter, a brief review has been conducted
on the activity models.

8.2 Water Content of Hydrocarbon Gases

8.2.1 Lean Sweet Natural Gas Water Content Correlation

The water vapor content of natural gases in equilibrium with water is commonly esti-
mated from Fig. 6.1 of Campbell book [1] or Fig. 20.4 of Gas Processors and Suppliers
Association, including corrections for the molecular weight (relative density) of gas
and salinity of water [2].
Figure 8.1a, b [3] present the saturated water content of lean sweet natural gasses.
Example 8.1 demonstrate the application of Fig. 8.1. The SI units of water content
are in kg of water per million cubic meters at standard conditions of 15 °C and
8 Water-Hydrocarbons System 353

a 100000
80000 0.7 MPa
60000
40000

3.5
20000
7.0
10000
8000
6000 17.5
4000 35.0
Water Content, kg water/106 Sm3 gas @ 15 °C and 101.3 kPa

52.5
2000
70.0 MPa

1000
800
600

400

200

100
80
60

40

20

10
8
6
4

1
-40 -30 -20 -10 0 10 20 30 40 50 60 70 80 90 100 110
Temperature, °C

Fig. 8.1 Saturated water content of sweet lean natural gas a SI unit b field unit [3]

101.3 kPa. The corresponding field units are lbm of water per million cubic foot at
standard conditions of 60 °F and 14.7 psia.

Example 8.1 Determine the saturated water content of a lean sweet natural
gas at 38 °C (100 °F) and 6.9 MPa (1000 psia) in kg/106 Sm3 (lbm/MMscf),
ppm, and mole fraction.
SI Solution
354 A. Izadpanahi and R. Azin

b 10000 100 psia

500
Water Content, lbm water/ MMSCF gas @ 60°F and 14.7 psia

1000 1 000

2 500
5 000
7 500

100 10 000 psia

10

0.1
-40 -20 0 20 40 60 80 100 120 140 160 180 200 220
Temperature, °F

Fig. 8.1 (continued)

From Fig. 8.1a, at 38 °C and 6.9 MPa, W = 990 kg/106 Sm3 .


y = (990 kg water/106 Sm3 ) (kmole/18 kg water) (23.64 Sm3 /kmole gas)
= 1300 ppm.
y = (1300 ppm) (1/106 ) = 0.0013 mol fraction.
Field Solution
From Fig. 8.1b, at 100 °F and 1000 psia, W = 61 lbm/MMscf.
y = (61 lbm water/MMscf) (lbmole/18 lbm water) (379.5 scf/lbmole gas)
= 1286 ppm.
y = (1286 ppm) (1/106 ) = 0.0013 mol fraction.
8 Water-Hydrocarbons System 355

Equations 8.1 and 8.2 show direct relationships between water content, W, and
mole fraction, y.
   
−6 kg kg
y = 1.3133 × 10 W 6 and W 6 = 761421(y) (8.1)
10 Sm3 10 Sm3
   
−6 lbm lbm
y = 21.083 × 10 W and W = 47430(y) (8.2)
MMscf MMscf

Example 8.2 Determine the saturated water content of a lean sweet natural
gas at 38 °C (100 °F) and 2.07 MPa (300 psia) in kg/106 Sm3 (lbm/MMscf).
Hint: Because water content varies inversely with pressure use the following
linear interpolation
  
−W P2
W P = W P1 1
− 1
+ W P1
P1 − P2
1 1 P P1
SI Solution
From Fig. 8.1a, at 38 °C and P1 = 0.7 MPa, W P1 = 7600 kg/106 Sm3 .
P2 = 3.5 MPa, WP2 = 1712 kg/106 Sm

W P = 7600−1712 1
− 0.7
1
+ 7600 = 2728 106kgSm3
0.7 − 3.5
1 1 2.07
Field Solution
From Fig. 8.1b, at 100 °F and P1 = 100 psia, W P1 = 475 lbm/MMscf.
P2 = 500 psia, W P2 = 107 lbm/MMscf
W P = 475−107
1
− 1
1
300
− 100
1
+ 475 = 168 MMscf
lbm
100 500

8.2.2 Propane Water Content Charts

For the binary propane–water system, the coexistence of equilibrium phases depends
on the system pressure and temperature as follows:
a. Propane vapor phase in equilibrium with liquid water
b. Propane vapor phase in equilibrium with ice or hydrate
c. Propane liquid phase in equilibrium with liquid water
d. Propane liquid phase in equilibrium with hydrate.
Figure 8.2 illustrates the presence of these equilibrium phases as a function of
temperature for the isobar of 14.7 psia (101.3 kPa).
Table 8.1 presents the estimated three phase temperatures and the saturation
temperatures of propane–water system.
The water content chart of propane vapor and liquid phases as a function of
temperature for six isobars are presented in Fig. 8.3. Because of the very small
values, the liquid propane water contents for different isobars fall on the same curve.
356 A. Izadpanahi and R. Azin

Temperature, °C
-51 -40 -29 -18 -7 4 16 27 38 49 60 71 82 93 104
1.0E+00 1.0E+00
Water mole fracƟon in propane vapor or liquid phase in equilibrium with liquid water, hydrate,

1.0E-01 1.0E-01

C3 Vapor phase
in equilibrium
1.0E-02 1.0E-02
with liquid water

1.0E-03 1.0E-03
or ice

1.0E-04 1.0E-04

Freeze temperature
C3 Vapor phase
1.0E-05 transiƟon to 1.0E-05
liqud phase

1.0E-06 1.0E-06
C3 Liquid phase
in equilibrium
with hydrate 14.7 psia (101.3 kPa)

1.0E-07 1.0E-07
-60 -40 -20 0 20 40 60 80 100 120 140 160 180 200 220
Temperature,° F

Fig. 8.2 Water content of vapor and liquid propane as a function of temperature at 14.7 psia
(101.3 kPa) [4]

Example 8.3 Determine the saturated water content of propane vapor at 38 °C


(100 °F) and 0.69 MPa (100 psia) in mole fraction using Figs. 8.1 and 8.3.
SI Solution
From Fig. 8.1a, at 38 °C and 0.69 MPa, W = 7600 kg/106 Sm3
8 Water-Hydrocarbons System 357

Table 8.1 Three phase estimated temperature and saturation temperature for six isobars [4]
Pressure 3 phase temperature Saturation temperature
psia kPa °F °C Phases present °F °C
14.7 101.4 27.5 −2.5 V–I–H −43.3 −41.8
25 172.4 32.1 0.1 V–LW –H −20.4 −29.1
50 344.8 37.9 3.3 V–LW –H 14.0 −10.0
100 689.7 41.8 5.4 LP –LW –H 54.8 12.7
150 1034.5 41.8 5.4 LP –LW –H 82.3 27.9
200 1379.3 41.8 5.4 LP –LW –H 103.7 39.8
Phases V Vapor, I Ice, H Hydrate, L W Liquid Water, and L P Liquid Propane

   
y = 1.3133 × 10−6 W 106kgSm3 = 1.3133 × 10−6 7600 106kgSm3 =
0.01 Eq. 8.1
From Fig. 8.3, at 38 °C and 0.69 MPa, y = 0.009.
Field Solution
 °Flbmand 100 psia, W = −6
From Fig. 8.1b, at 100 480 lbm/MMscf

y = 21.083 × 10−6 W MMscf = 21.083 × 10 480 MMscf
lbm
= 0.01 Eq. 8.2
From Fig. 8.3, at 100 °F and 100 psia, y = 0.009.

Example 8.4 Determine the saturated water content of propane liquid at 10 °C


(50 °F) and 0.69 MPa (100 psia) in mole fraction using Fig. 8.3.
SI Solution
From Fig. 8.3 at 38 °C and 0.69 MPa, x = 0.00015 liquid water mole
fraction.
Field Solution
From Fig. 8.3, at 50 °F and 100 psia, x = 0.00015 liquid water mole fraction.

Estimating propane water content requires a good understanding of the phase


behavior. The process simulation programs like ProMax [5] have several tools or
procedures for estimating the water content. Which one should be used to give a
correct answer? In addition to the selection of a suitable equation of state, the selection
of the right tool or procedure at a given set of conditions is essential. The choice of
a suitable tool changes as the conditions or the equilibrium phases change.
The presented propane water content chart can be used for facility type calcula-
tions and trouble shooting. It is a good practice to test the performance/accuracy of
the selected tool against experimental data first. Obviously, for better understanding
of propane–water phase behavior and improving the thermodynamic modeling more
experimental data are needed.
358 A. Izadpanahi and R. Azin

Temperature, °C

-51 -40 -29 -18 -7 4 16 27 38 49 60 71 82 93 104


1.0E+00 1.0E+00
Water mole fracƟon in propane vapor or liquid phase in equilibrium with liquid water, hydrate,

Freeze
temperature
1.0E-01 1.0E-01

1.0E-02 1.0E-02

Vapor-Liquid
Phase TransiƟon
1.0E-03 1.0E-03
or ice

1.0E-04 1.0E-04

1.0E-05 1.0E-05

14.7 psia (101.3 kPa)


25 psia (172 kPa)
1.0E-06 50 psia (345 kPa) 1.0E-06
100 psia (699 kPa)
150 psia (1034 kPa)
200 psia (1379 kPa)
1.0E-07 1.0E-07
-60 -40 -20 0 20 40 60 80 100 120 140 160 180 200 220
Temperature,° F

Fig. 8.3 Water content of propane as a function of temperature for six isobars [4]

8.2.3 Ethane Water Content Charts

The coexistence of equilibrium phases depends on the system pressure and


temperature as follows:
e. Ethane vapor phase in equilibrium with liquid water
8 Water-Hydrocarbons System 359

f. Ethane vapor phase in equilibrium with ice or hydrate


g. Ethane liquid phase in equilibrium with liquid water
h. Ethane liquid phase in equilibrium with hydrate.
Table 8.2 presents the three phase temperatures of a ethane–water system and the
saturation temperatures of pure ethane [6].
Figure 8.4 illustrates the presence of these equilibrium phases as a function of
temperature for the isobar of 200 psia (1379 kPa) [6].
Similarly, the water content charts of ethane vapor and liquid phases for the other
isobars are presented in Fig. 8.5. Due to the very small values the liquid ethane water
contents for different isobars fall on the same curve.

Example 8.5 Determine the saturated water content of ethane vapor at 38 °C


(100 °F) and 0.69 MPa (100 psia) in mole fraction using Figs. 8.1 and 8.5.
SI Solution
 and 0.69MPa, W = 7600 kg/10 Sm
6 3
From Fig. 8.1a, at 38 °C 
y = 1.3133 × 10−6 W 106kgSm3 = 1.3133 × 10−6 7600 106kgSm3 =
0.01 Eq. 8.1
From Fig. 8.5, at 38 °C and 0.69 MPa, y = 0.0095.
Field Solution
 °Flbmand 100 psia, W = −6
From Fig. 8.1b, at 100 480 lbm/MMscf
y = 21.083 × 10−6 W MMscf = 21.083 × 10 480 MMscf
lbm
= 0.01 Eq. 8.2
From Fig. 8.5, at 100 °F and 100 psia, y = 0.0095.

Table 8.2 Three phase temperature and saturation temperature for six isobars
Pressure 3 phase temperature Saturation temperature
psia kPa °F °C Phases present °F °C
14.7 101.4 32 0 V–LW –I −127.0 −41.8
25 172.4 32 0 V–LW –I −108.2 −29.1
50 344.8 27.5 −2.5 V–LW –H −80.0 −10.0
100 689.7 37.5 3.1 V–LW –H −46.5 12.7
150 1034.5 43.5 6.4 V–LW –H −23.9 27.9
200 1379.3 47.5 8.6 V–LW –H −6.3 39.8
Phases V Vapor, I Ice, H Hydrate, and L W Liquid Water
360 A. Izadpanahi and R. Azin

Temperature, °C

-51 -40 -29 -18 -7 4 16 27 38 49 60 71 82 93 104


1.0E-01 1.0E-01
Water mole fracƟon in ethane vapor or liquid phase in equilibrium with liquid water, hydrate

200 psia (1379 kPa)

1.0E-02 C2 Vapor phase in equilibrium 1.0E-02


with liquid water

1.0E-03 1.0E-03

C2 Vapor phase in equilibrium


or ice

1.0E-04 with hydrate 1.0E-04

C2 Vapor phase transiƟon to


liqud phase
1.0E-05 1.0E-05

Freeze temperature

1.0E-06 1.0E-06
C2 Liquid phase in equilibrium
with hydrate

1.0E-07 1.0E-07
-60 -40 -20 0 20 40 60 80 100 120 140 160 180 200 220
Temperature,° F

Fig. 8.4 Water content of vapor and liquid Ethane as a function of temperature at 200 psia (1379 kPa)
[6]

8.2.4 Acid Gases Water Content Charts

The phase equilibria in the system H2 S + water and CO2 + water are key to the
discussion of the water content of an acid gas system. Figure 8.6a (SI) and 8.6b (Field)
present the water content of pure H2 S as a function of pressure and temperature [7].
The behavior shown on this plot is quite complicated and explained thoroughly by
Carroll [8]: “At low pressure the hydrogen sulfide + water mixture is in the gas
8 Water-Hydrocarbons System 361

Temperature, °C

-51 -40 -29 -18 -7 4 16 27 38 49 60 71 82 93 104


1.0E+00 1.0E+00
Water mole fracƟon in ethane vapor or liquid phase in equilibrium with liquid water, hydrate

1.0E-01 1.0E-01

1.0E-02 1.0E-02

1.0E-03 1.0E-03
or ice

1.0E-04 Freeze temperature 1.0E-04

1.0E-05 1.0E-05
14.7 psia (101.3 kPa)
25 psia (172 kPa)
50 psia (345 kPa)
1.0E-06 1.0E-06
100 psia (699 kPa)
150 psia (1034 kPa)
200 psia (1379 kPa)
1.0E-07 1.0E-07
-60 -40 -20 0 20 40 60 80 100 120 140 160 180 200 220
Temperature,° F

Fig. 8.5 Water content of Ethane as a function of temperature for six isobars

phase. At low pressure the water content tends to decrease with increasing pressure,
which is as expected. Eventually a pressure is reached where the H2 S is liquefied.
This plot is represented by the discontinuity in the curve and a broken line joins
the phase transition. There is a step change in the water content when there is a
transition from vapor to liquid. In the case of hydrogen sulfide, the water content of
the H2 S liquid is greater than the coexisting vapor. This is contrary to the behavior
362 A. Izadpanahi and R. Azin

for light hydrocarbons where the water content in the hydrocarbon liquid is less than
the coexisting vapor.”
Within the transition region, the acid gas exists as both liquid and vapor. Water
saturation of the vapor phase is represented by the lower value, whereas the water
content of the liquid phase is the higher value.
Figure 8.7a, b present the water content of pure CO2 and pure CH4 as a func-
tion of pressure and temperature. When Figs. 8.6 and 8.7 agree very good with the
experimental data and the Yarrison et al. model [9].

Example 8.6 Determine the saturated water content of H2 S vapor at 38 °C


(100 °F) and 2.07 MPa (300 psia) in kg/106 Sm3 (lbm/MMscf) using Fig. 8.6
and compare it with the water content of sweet gas of Example 8.2.
SI Solution
From Fig. 8.6a, at 38 °C and 2.07 MPa, WH2 S = 2970 kg/106 Sm3 .
From Fig. 8.1a, and Example 8.1 at 38 °C and 2.07 MPa, W Sweet Gas =
2728 kg/106 Sm3 .
Field Solution
From Fig. 8.6b, at 100 °F and 300 psia, WH2 S = 185 lbm/MMscf.
From Fig. 8.1b, at 100 °F and 100 psia, W Sweet Gas = 168 lbm/MMscf.
In general, acid gases like H2 S and CO2 and Sour gases can hold more water.

In general the phase behavior of the system CO2 + water is as complex as that
of the H2 S + water system. The CO2 -rich liquid phase only occurs for temperatures
less than about 32.2 °C (90 °F). As shown in Fig. 8.7, the water content of CO2
exhibits a minimum.
Notice Fig. 8.7a, b indicate that at the same pressure and temperature CO2 can
hold more water than CH4 .
Figure 8.8 presents the phase behavior of pure CO2 , Pure H2 S and three mixtures
containing 2 mol percent CH4 . Their corresponding water content charts are presented
in Fig. 8.9.

8.2.5 Sour Natural Gas Water Content Charts

There are several methods of calculating of water content of sour gases. and the modi-
fied Soave–Redlich–Kwong (SRK EoS) reported in GPA RR-42 by Erbar et al. [10].
This version of SRK is tailor-fitted to handle water-hydrocarbon systems containing
hydrogen sulfide and carbon dioxide.
8 Water-Hydrocarbons System 363

a 1000000
800000
600000

400000

200000
Water content, kg/106 sm3

100000 171.1C
80000
60000 137.8C
40000
104.4C

20000 71.1C

10000 37.8C
8000
6000
4000

2000

1000
100.00 1000.00 10000.00 100000.00
Pressure, kPa

b 100000
80000
60000
40000

20000
Water content, lb/MMSCF

10000
8000
340F
6000
4000 280F

220F
2000

160
1000
800 100F
600
400

200

100
100 1000 10000

Pressure, psia

Fig. 8.6 Isothermal water content of pure H2 S [7]


364 A. Izadpanahi and R. Azin

a 100000
80000
60000
40000
75C CO2
20000 50C CO2
31.1C CO2
Water content, kg/106 Sm3

10000 25C CO2


8000
18.3C CO2
6000
18.3C CH4
4000
31.1C CH4
50C CH4
2000
75C CH4

1000
800
600

400

200

100
100 1000 10000
Pressure, kPa
b 4000
3000

2000
167F CO2
122F CO2
1000
800 88F CO2
Water content, lb/MMSCF

600
77F CO2
400
65F CO2

65F CH4
200
88F CH4

122F CH4
100
80 167F CH4
60

40

20

10
100 1000 10000
Pressure, psia

Fig. 8.7 Isothermal water content of pure CO2 and CH4 [7]
8 Water-Hydrocarbons System 365

Temperature, °F
-58 -40 -22 -4 14 32 50 68 86 104 122 140 158 176 194 212 230
10000 1450
37.8 C (100 F) Isotherm
9000 2-25-73 (C1-H2S-CO2) 1305
2-49-49 (C1-H2S-CO2)
8000 2-75-23 (C1-H2S-CO2) 1160
Pure CO2
7000 Pure H2S 1015

Pressure, psia
Pressure, kPaa

6000 870

5000 725

4000 580

3000 435

2000 290

1000 145

0 0
-50 -40 -30 -20 -10 0 10 20 30 40 50 60 70 80 90 100 110
Temperature, °C

Fig. 8.8 Phase behavior of pure CO2 , pure H2 S and their mixtures with 2 mol% CH4 [7]

8.2.5.1 Maddox et al.

In this method water content of sour natural gas is estimated by Eq. 8.3 [11].

WSour = ysweet Wsweet + yCO2 WCO2 + yH2 s WH2 s (8.3)

where:
W Sour water content of gas containing H2 S and/or CO2 .
W Sweet water content of sweet part of gas from Fig. 8.1
WCO2 effective water content of CO2 from Fig. 8.10a,
WH2 S effective water content of H2 S from Fig. 8.11a, b
ySweet ol fraction of hydrocarbons.
yCO2 mole fraction of CO2 .
yH2 S mole fraction of H2 S.
Figures 8.10 and 8.11 show what is called the “effective water content” of the
acid gas component. This represents the effective water content of the component
in the natural gas mixture, not the water content of the pure component. These
366 A. Izadpanahi and R. Azin

a 20000
100% CO2
25% H2S+73% CO2+2% CH4
49% H2S+49% CO2+2% CH4
10000 75% H2S+23% CO2+2% CH4
Water content, kg/106 Sm3

8000 100% H2S


6000

4000

2000

1000
100.00 1000.00 10000.00

Pressure, kPa

b 1000
100% CO2
8000
25% H2S+73% CO2+2%
CH4
6000
49% H2S+49% CO2+2%
Water content, lb/MMSCF

CH4

4000

2000

100
100 1000 3000

Pressure, psia

Fig. 8.9 Water content of mixtures of 2 mol% CH4 , H2 S and CO2 , at 37.8 ◦ C a SI unit b Field
unit [7]

curves were based on the experimental data Sharma. Figures 8.10 and 8.11 were
obtained by cross-plotting and smoothing Sharma’s binary data for methane-CO2 ,
and methane-H2 S [11].
8 Water-Hydrocarbons System 367

Fig. 8.10 Water content contribution of CO2


368 A. Izadpanahi and R. Azin

Fig. 8.11 Water content contribution of H2 S


8 Water-Hydrocarbons System 369

Example 8.7 Determine the saturated water content of a sour natural gas at
71 °C (160 °F) and 6.9 MPa (1000 psia) in kg/106 Sm3 (lbm/MMscf), for the
following cases:
a. containing 0.20 mol fraction CO2
b. containing 0.17 mol fraction H2 S
c. containing 0.20 mol fraction CO2 and 0.17 mol fraction H2 S.
SI Solution
From Fig. 8.1a, at 71 °C and 6.9 MPa, W Sweet = 4360 kg/106 Sm3 .
From Fig. 8.10a, at 71 °C and 6.9 MPa, WCO2 = 4700 kg/106 Sm3 , Using
Eq. 8.2
a. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.8(4360) +
0.2(4700) = 4428 kg/106 Sm3
From Fig. 8.11a, at 71 °C and 6.9 MPa, WH2 S = 5900 kg/106 Sm3 , Using
Eq. 8.2
b. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.83(4360) +
0.17(5900) = 4622 kg/106 Sm3
c. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.63(4360) +
0.2(4700) + 0.17(5900) = 4690 kg/106 Sm3
Field Solution
From Fig. 8.1b, at 160 °F and 1000 psia, W Sweet = 275 lbm/MMscf.
From Fig. 8.10b, at 160 °F and 1000 psia, WCO2 = 290 lbm/MMscf Using
Eq. 8.2
a. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.8(275) + 0.2(290) =
278 lbm water/MMscf
From Fig. 8.11b, at 160 °F and 1000 psia, WH2 S = 380 lbm/MMscf Using
Eq. 8.2
b. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.83(275) +
0.17(380) = 293 lbm water/MMscf
c. WSour = ySweet WSweet + yCO2 WCO2 + yH2 S WH2 S = 0.63(275)+0.2(290)+
0.17(380) = 296 lbm water/MMscf

8.2.5.2 Wichert and Wichert Chart

Based on the Wichert and Wichert chart [12], Fig. 8.12a, b, the sour gas water content
is estimated by multiplying the sweet gas water content by a sourness correction
factor, F. This correction factor, F, is a function of H2 S equivalent concentration
(HEC), temperature and pressure. The H2 S equivalent concentration is defined by
370 A. Izadpanahi and R. Azin

Fig. 8.12 Sour gas: sweet gas–water content ratio [12]


8 Water-Hydrocarbons System 371

Eq. 8.4.

HEC = Mole%H2 S + 0.7(Mole%CO2 ) (8.4)

To use Fig. 8.12a, b, enter the lower portion of the graph at the system temperature.
Proceed.
horizontally to the equivalent H2 S concentration. Proceed vertically to the system
pressure then horizontally to the y-axis. The “water content ratio” factor should be
multiplied by the sweet water content (at the same T & P) to obtain the water content
of the sour gas.

Example 8.8 Determine the saturated water content of a sour natural gas at
71 °C (160 °F) and 6.9 MPa (1000 psia) in kg/106 Sm3 (lbm/MMscf), if the
gas contains 0.20 mol fraction CO2 and 0.17 mol fraction H2 S.
SI Solution
From Fig. 8.1a, at 71 °C and 6.9 MPa, W Sweet = 4360 kg/106 Sm3 .
H2 S equivalent of CO2 , HEC, using Eq. 8.4
HEC = 0.70yCO2 + yH2 S = 0.7(0.2) + 017 = 0.31 = 31%
From Fig. 8.12a, at 71 °C, HEC = 31% and 6.9 MPa, F= Correction Factor
= 1.1
WSour = (Correction Factor)WSweet = 1.1(4360) = 4796 kg/106 Sm3
Field Solution
From Fig. 8.1b, at 160 °F and 1000 psia, W Sweet = 275 lbm/MMscf.
H2 S equivalent of CO2 , HEC, using Eq. 8.4
HEC = 0.70yCO2 + yH2 S = 0.7(0.2) + 017 = 0.31 = 31%
From Fig. 8.12b, at 160 °F, HEC = 31% and 1000 psia, = Correction Factor
= 1.1
WSour = (Correction Factor)WSweet = 1.1(275) = 303 lbm water/MMscf.

Table 8.3 indicates that as long as the total acid gas concentrations is less than
60 mol percent, Maddox et al. and the modified SRK methods produce results within
the accuracy of experimental data. However, for higher concentrations of acid gases,
the modified SRK provides a better prediction.
The composition, experimental, and predicted water content by Maddox et al.
and the modified SRK for two natural gas mixtures are presented in Table 8.4. The
upper part of columns 1, 2, 4, 5 report the measured mole percent. Based on the
feed compositions shown in columns 1 and 4, three-phase flash calculations using
the modified SRK were performed and the resulting vapor stream compositions are
shown in columns 3 and 6, respectively. Notice that the measured and predicted vapor
compositions are not identical. Inaccuracies in predicting the vapor composition can
result in errors in predicting the water saturation.
For each vapor stream, the saturated water content was predicted by both methods
and is presented in the lower portion of this table. As can be seen from this table, both
372 A. Izadpanahi and R. Azin

Table 8.3 Comparison of Maddox et al. [11] and modified SRK EOS results [10] with the
experimental water content [13] of several sour gas mixtures
Mole % T, °F/°C P, Water content, lbm/MMSCF/kg/106 std m3
CO2 H2 S CH4 psia/MPa Experimental Maddox et al. SRK
11.0 – 89.0 100/37.8 2000/13.79 40.6/652 40.7/653 40.5/650
11.0 – 89.0 160/71.1 1000/6.89 286.0/4591 272.4/4372 283.3/4548
20.0 – 80.0 100/37.8 2000/13.79 40.6/652 42.8/688 44.8/719
20.0 – 80.0 160/71.1 1000/6.89 282.0/4527 273.3/4387 293.1/4704
– 8.0 92.0 130/54.4 1500/10.34 111.0/1782 108.3/1739 103.6/1664
– 27.5 72.5 160/71.1 1367/9.43 247.0/3965 257.7/4136 260.9/4188
– 17.0 83.0 160/71.1 1000/6.89 292.0/4628 288.3/4628 297.5/4776
60.0 10.0 30.0 100/37.8 1100/7.58 81.0/1300 69.7/1118 84.5/1357
10.0 81.0 9.0 100/37.8 1900/13.10 442.0/7095 87.7/1408 360.0/5779
94.7 – 5.3 77/25 1500/10.34 109.2/1753 44.8/720 137.8/2212
94.7 – 5.3 122/50 2000/13.79 164.6/2642 105.3/1690 258.3/4147

Table 8.4 Conditions and compositions of two sour natural gases and their saturated water contents
[14]
Component 1 2 3 4 5 6
Mixture A Mole % Mixture B Mole %
Experimental SRK Experimental SRK
Feed Vapor Vapor Feed Vapor Vapor
CO2 0.5797 15.6448 16.766 3.0532 16.8993 17.089
H2 S 0.1936 3.7581 6.053 1.0198 4.6664 6.130
CH4 2.1369 72.1444 68.711 11.2540 70.4544 68.981
C2 H6 0.1356 4.4428 4.372 0.7141 4.4917 4.387
C3 H8 0.0623 1.9831 1.996 0.3283 2.0703 2.020
iC4 H10 0.0101 0.3165 0.322 0.0534 0.3374 0.327
nC4 H10 0.0269 0.8361 0.869 0.1415 0.8935 0.875
H2 O 96.8549 0.8740 0.911 83.4357 0.1871 0.190
Total 100 100 100 100 100 100
T, °F [°C] 120 120 120 120 120 120
[48.9] [48.9] [48.9] [48.9] [48.9] [48.9]
P, psia [kPa] 200 200 200 1500 1500 1500
[1379] [1379] [1379] [10,342] [10,342] [10,342]
Experimental Water content [15] 414.5 88.7
lbm/MMSCF [kg/106 std m3 ] [6654] [1424]
Maddox et al. Water content [11] 405.9 401.1 88.9 90.2
lb/MMSCF [kg/106 std m3 ] [6516] [6438] [1427] [1448]
SRK Water content [10] 430.7 432.1 88.9 90.4
lbm/MMSCF [kg/106 std m3 ] [6914] [6937] [1427] [1451]
8 Water-Hydrocarbons System 373

Temperature, C
-129 -101 -73 -46 -18 10 38 65 93
1400 BP 9.66
DP
1200 Hydrate Curve 8.28
Hyd 25% MeOH
Retrograde
1000 WDP Chart 6.90

Pressure, MPa
Pressure, psia

WDP SRK
800 5.52

600 4.14

400 2.76

200 1.38

0 0.00
-200 -150 -100 -50 0 50 100 150 200
Temperature, F

Fig. 8.13 Water-sour natural gas phase behavior for mixture A [14]

methods predict saturated water content reasonably well. Not surprisingly, the accu-
racy of both methods is improved slightly when the experimental vapor composition
is used rather than the predicted vapor composition.
Figure 8.13 represents the phase behavior for mixture A. This figure includes from
right to left: the water dew point, hydrate formation, 25 weight percent methanol
(MeOH) inhibited hydrate formation, hydrocarbon dew point, retrograde, and the
bubble point curves, respectively. The blue-triangular-symbol water dew point curve
is predicted by Maddox et al. method. The red curve represents the water dew point
predicted by rigorous calculations using the modified SRK [10]. It is interesting to
see that both methods agree quite well with each other. The region to the right of the
water dew point curve is gas phase and to the left, liquid water is present.
Figure 8.14 presents the phase behavior of sour gas mixture B. With exception of
low pressure region, both methods agree quite well.
Figure 8.15 demonstrates the effect of acid gases on phase behavior of mixture B.
As shown in this figure, the presence of acid gases shifts all of the curves to the right. In
other words, the presence of acid gases increases the hydrate formation temperature
considerably. It also increases the water dew point temperature. It should be noted that
the water dew point curves have been generated for a fixed amount of water content
predicted at specified separator condition. In this case the separator temperature and
pressure were 120 °F [48.9 °C] and 1500 psia [10,342 kPa], respectively.
374 A. Izadpanahi and R. Azin

Temperature, C
-129 -101 -73 -46 -18 10 38 65
1400 BP 9.65
DP
1200 Hydrate Curve 8.27
Hyd 25% MeOH
Retrograde
1000 WDP Chart 6.90

Pressure, MPa
Pressure, psia

WDP SRK
800 5.52

600 4.14

400 2.76

200 1.38

0 0.00
-200 -150 -100 -50 0 50 100 150
Temperature, F

Fig. 8.14 Water-sour natural gas phase behavior for mixture B [14]

Temperature, C
-129 -101 -73 -46 -18 10 38 65
1400 Sour BP
9.66
Sour DP
1200 Sour Hydrate 8.28
Sweet BP
Sweet DP
1000 Sweet Hydrate 6.90
Sour WDP

Pressure, MPa
Pressure, psia

Sweet WDP
800 5.52

600 4.14

400 2.76

200 1.38

0 0.00
-200 -150 -100 -50 0 50 100 150
Temperature, F

Fig. 8.15 Effect of acid gases on phase behavior for mixture B [14]
8 Water-Hydrocarbons System 375

8.3 Gas Solubility

8.3.1 Solubility of Hydrocarbons in Water

The solubility of different hydrocarbon components are discussed in this section.


This subject is very important due to the presence of water in hydrocarbon systems
including oil and gas reservoirs. Many researchers have studied the solubility of pure
and mixtures of hydrocarbon components in water at different temperatures and
pressures. Culberson and Mcketta (1950) reported the solubility of ethane in water
for different temperatures and pressures. Figure 8.16a shows the reported values of
isotherm solubility of ethane in water at low pressures. As it is obvious from this
figure, the solubility of ethane increases by increasing the pressure. However, the
solubility at constant pressure decreases until a specific temperature (Fig. 8.16b).
After that temperature, the solubility increases by increasing the temperature. This
phenomenon is called the minimum solubility [16].
The solubility of ethane in high pressure was reported in another study by
Culberson and Mcketta (1950). Figure 8.17a indicates the solubility of ethane at
pressures up to 10,000 psi for different temperatures. As it is shown, at very high
pressure and low temperature, the solubility behavior of ethane in water is complex.
But, the solubility is increased by increasing the pressure. Figure 8.17b shows the
isobar solubility of ethane in the water. The solubility of ethane in water increases
by increasing the temperature at 10,000 psi. So, minimum solubility of ethane which
expressed earlier is vanished at the mentioned pressure for temperatures below 340
°F. But, the minimum solubility is seen for pressures below 10,000 psi [17].
Culberson and Mcketta (1951) studied the solubility of methane at the same
conditions as their second study. Figure 8.18a show the solubility of methane in water
at different temperatures and pressures. The solubility of methane also increases by
increasing the temperature. Figure 8.18b shows the isobar solubility of methane.
The minimum solubilities happen in the same range of temperatures for the methane-
water and ethane water systems. This would not necessarily have been expected, since
the temperatures are very near to the critical of ethane and considerably removed
from that of methane. Evidently the critical properties of the solute do not constitute
a good basis for correlating solubility. Regarding the ethane-water system, for the
higher pressures the methane water system also shows the probability of a maximum
solubility following the minimum solubility as the temperature is increased [18].
They also compared the solubility of methane, ethane and natural gas for different
conditions. Figure 8.19 shows the solubility behavior of three mentioned gases in
water. This figure indicates that solubilities of the ethane and the methane are very
similar to each other. Also, the solubility of the methane being slightly bigger at 500
psia. As the pressure increases, the solubility of the methane increases more than that
of the ethane. The solubility of the methane is more than twice that of the ethane at
10,000 psia. However, the solubility of the ethane is 30 percent bigger than that of the
methane at one atmosphere. The natural gas–water system contains 88.51% methane
and 6.02% ethane. The solubility of the natural gas is greater than that of the pure
376 A. Izadpanahi and R. Azin

1210 (a)

1010
Pressure (psi)

810

610

100 F
410 160 F
220 F

210 280 F
340 F

10
0 0.0002 0.0004 0.0006 0.0008 0.001
Mole Fraction Ethane
0.0014
(b) 1200 psi
900 psi
0.0012 700 psi
Mole Fraction Ethane

500 psi
300 psi
0.001
100 psi

0.0008

0.0006

0.0004

0.0002

0
100 150 200 250 300
Temperature (F)

Fig. 8.16 Solubility of ethane in water in low pressure a isotherm solubility b isobar solubility
[16]

methane at the lower pressures. Since the ethane present and the other hydrocarbons
maybe, nearly as soluble as the methane at low pressures. The presence of these
components increases the solubility of the natural gas to a value bigger than pure
methane in low pressures. Their relatively small solubility in comparison to methane
at the higher pressures reduces the solubility of the natural gas below the solubility
of pure methane [18].
Reamer et al. (1952) studied the behavior of n-butane-water system. Figure 8.20
shows the mole fraction of n-butane in aqueous liquid phase (water) at different
8 Water-Hydrocarbons System 377

10000
(a)
9000

8000

7000
Pressure (psi)

6000

5000

4000
100 F

3000 160 F
220 F
2000
280 F
1000 340 F

0
0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035

Mole Fraction Ethane


0.0035
(b) 10000 psi
7000 psi
0.003 5000 psi
Mole Fraction Ethane

3000 psi
2000 psi
1500 psi
0.0025

0.002

0.0015

0.001

0.0005
100 150 200 250 300 350

Temperature (F)

Fig. 8.17 Solubility of ethane in water at high pressures a isotherm solubility b isobar solubility
[17]

conditions. This figure indicates that the mole fraction of n-butane in the water
increases by increasing the temperature at high pressure conditions. However, this
effect is reversed at the lower pressures. They also stated that the behavior of n-
butane-water system at low pressures and temperatures was not established with
certainty [19].
Song et al. (1996) studied the solubility of pure ethane and methane under different
conditions including the hydrate formation points. Their experiments were conducted
in the following conditions: (1) methane at 3.45 MPa and temperatures ranging from
378 A. Izadpanahi and R. Azin

10000
(a)
9000

8000

7000
Pressure (psi)

6000

5000

4000
77 F
3000 100 F
160 F
2000 220 F
280 F
1000 340 F

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008

Mole Fraction Ethane


0.008
(b) 10000 psi
6000 psi
0.007
3000 psi
1500 psi
Mole Fraction Ethane

0.006 600 psi


200 psi
0.005

0.004

0.003

0.002

0.001

0
60 110 160 210 260 310 360

Temperature (F)

Fig. 8.18 Solubility of methane in water a isotherm solubility b isobar solubility [18]

290.2 to 273.2 K, (2) ethane at 0.66 MPa (and a temperature range of 290.2 to
273.2 K. They reported that at the lower temperatures the isobaric solubility showed
a significant difference from Henry’s law solubility. The trapping of gas molecules in
water structure is caused by the increasing in solubility. Also, they observed several
sub-processes as the hydrate is formed including: (1) the dissolution of the gas in the
liquid water, or interstitial solubility, (2) the onset of the sorption sub-process and
build-up of the hydrate precursors, (3) the catastrophic formation and the existence
8 Water-Hydrocarbons System 379

0.008
10000 psi
Methane

0.007 Ethane

Natural gas
0.006
Mole Fraction

0.005

0.004

5000 psi
0.003 5000 psi 10000 psi

3000 psi 3000 psi


0.002

0.001
500 psi 500 psi
500 psi
0
80 130 180 230 280 330 380
Temperature (F)

Fig. 8.19 Solubility behavior of methane, ethane, and natural gas in water [18]

of the solid phase with its slushy characteristics, and (4) the solidification of the
hydrates which causes the total plugging of the experimental cell [20].
Dhima et al. (1998) reported the solubility data of methane, ethane and n-butane
at the 344.25 K and the pressures between 2.5–100 MPa. Their experimental results
categorize as pure hydrocarbon-water, methane-ethane-water, methane-n-butane-
water, and methane-n-pentane-water. They developed a model to predict the solubility
of hydrocarbons base on the Peng-Robinson equation of state modified by Soreide
and Whitson. They also predicted the solubility of a natural gas composition in water
based on their developed model. Their results show that the solubility of methane
in binary mixture of methane- water system is higher than the solubility of methane
in ternary methane-n-pentane-water. Also, the difference between the solubility of
methane increases by increasing the system temperature. This fact is shown in Fig. 12
of Dhima et al. [21].
Chapoy et al. (2004a) studied the solubility of methane-ethane-n-butane mixture
in water in different pressures and low temperatures (278.14–313.12 K). Studied
380 A. Izadpanahi and R. Azin

0.002

0.0018

0.0016

0.0014 10000 psi


Mole Fraction n-Butane

5000 psi
0.0012
1000 psi

0.001 500 psi


200 psi
0.0008

0.0006

0.0004

0.0002

0
100 150 200 250 300 350 400 450 500
Temperature (F)

Fig. 8.20 Solubility behavior of n-butane in water [19]

mixture contained 94%, 4% and 2% of the methane, ethane and n-butane, respec-
tively. Table 8.5 shows the experimental mole fractions of methane, ethane and
n-butane (x1 , x2 and x3 , respectively). This table indicates that at constant temper-
ature, x1 , x2 and x3 increase by increasing the pressure. Also, it can be understood
at constant pressure the mole fraction of n-butane is higher than that for ethane and
methane [22] (Table 8.5).
Chapoy et al. (2004b) reported the experimental solubility data of propane in
water for the temperatures between (277.62–368.16 °K) and the pressures to 568
psi. Figure 8.21 shows their experimental data. Also, they modelled the solubility
of the propane in water using the Valderrama modification of the Patel–Teja equa-
tion of state. The results showed that the thermodynamic model could predict the
experimental data with high accuracy [23].
Reshadi et al. (2012) presented a new mixing rule for accurate prediction of the
natural gas solubility in liquid water phase. They used the literature experimental
data of binary mixtures to tune the parameters. The results of their study show that
the predicted values of solubility are in good agreement with experimental data [24].
Mohebbi et al. (2012) reported the solubility of i-butane and the henry’s law
constant for light hydrocarbons in the water at low temperature. Also, they developed
a model based on Krichevsky–Kasarnovsky equation to consider the temperature and
pressure effects on the equation parameter. Their results show that temperature has
a significant effect on the Henry’s constant of the selected components. They also
8 Water-Hydrocarbons System 381

Table 8.5 The experimental solubility data of methane, ethane, and n-butane in water reported by
Chapoy et al. [22]
T(K) P (psi) Methane mole fraction Ethane mole fraction n-Butane mole fraction
×10e−5
278.14 149.679 0.000339 0.0000199 0.703
278.15 290.656 0.000646 0.0000387 1.121
283.14 143.152 0.000295 0.0000172 0.512
283.16 150.549 0.0003 0.0000165 0.482
283.16 288.335 0.000566 0.0000285 0.754
283.15 301.243 0.000593 0.0000302 0.941
283.15 446.571 0.000826 0.0000385 1.063
283.14 495.014 0.000891 0.0000428 1.121
283.15 495.304 0.000896 0.0000436 1.048
288.16 150.549 0.000282 0.0000172 0.577
288.17 444.976 0.000755 0.0000399 0.734
298.14 144.167 0.000218 0.0000147 0.387
298.14 429.892 0.000637 0.000033 0.811
298.14 1052.54 0.001359 0.0000562 0.991
298.14 1704.05 0.002014 0.0000674 –
298.14 2089.56 0.002191 0.0000672 –
313.12 1081.98 0.001157 0.0000406 0.694
313.12 1830.96 0.001817 0.0000586 –

0.0003

0.00025
Mole Fraction Propane

0.0002
278.09 K
280.14 K
0.00015 283.06 K
288.13 K
293.13 K
0.0001 298.12 K
308.13 K
323.13 K
0.00005 338.15 K
353.18 K
368.16K
0
0 100 200 300 400 500 600
Pressure (psi)

Fig. 8.21 Experimental solubility of propane in water [23]


382 A. Izadpanahi and R. Azin

reported that the effect of pressure in the Henry’s constant is negligible. However,
this effect in the components like methane, ethane and propane is considerable [25].

8.3.2 Solubility of Non-hydrocarbon Gases in Water

The behavior of non-hydrocarbon gases in contact with water is very important due
to their existence in the hydrocarbon reservoirs. These gases are naturally existed
in the composition of the fluids in the hydrocarbon reservoirs. Also, these gases are
injected into the geological formations in processes such as pressure maintenance,
CO2 storage, enhanced oil and gas recovery and acid gas disposal. As the geological
formations contain water, the solubility of these gases plays an important role in
the planning of each mentioned processes. In this section, the studies related to
determining CO2 , H2 S and N2 solubilities in water are investigated.
Wiebe and Gaddy (1939) reported a procedure for measuring the solubility of
carbon dioxide (CO2 ) in water. They reported the solubility data at 50, 75 and 100 °C
from 25 to 700 atm. Figure 8.22a, b show the solubility of CO2 at different pres-
sures and constant temperature and different temperatures and constant pressure,
respectively [26].
Wiebe and Gaddy (1940) determined the solubility of CO2 in water at the pressures
to 500 atmospheres and temperature lower than their previous work (between 12 to
40 °C). They observed complete miscibility at low temperature rate. Also, they
showed the critical curve qualitatively. They reported that the solid carbon dioxide
hexahydrate was formed in the following ranges: (1) 10 °C—all pressures above
50 atm (2) 12 °C—between 300 and 500 atm. Figure 8.23a shows the mole fraction
of CO2 in water at various temperature. Also, the critical curve is shown qualitatively.
They stated that their setup was not able to determine the change of critical solution
pressure with either composition or temperature. Figure 8.23b indicates the solubility
isotherms. Also, this figure illustrates the change of state of CO2 from gas to liquid
[27].
Yang et al. (2000) reported an apparatus for determination of the dissociation
pressure of hydrate-liquid water–vapor CO2 and hydrate-liquid water–liquid CO2 .
Also, they measured the CO2 solubility in liquid water. They reported the solubility
data in Table 8.4 of their work [28].
Chapoy et al. (2004) reported solubility data of CO2 in water in a wide temperature
and pressure range. Their experimental method is based on the measurement of
bubble-point pressure at isothermal conditions using a variable volume PVT cell.
They also validated their procedure by comparing their results with the results that
published in the literature. Table 8.6 shows the experimental data reported by Chapoy
et al. [29].
Wright and Maass (1932) studied the solubility of H2 S in water between 5 and
60 °C. They reported the experimental data of solubility. They expressed that by
increasing pressure at constant temperature the moles of H2 S in the solution will be
8 Water-Hydrocarbons System 383

40
(a)
35
Solubility (cc/grWater)

30

25

20

50 C
15
75 C

10 100 C

0
0 100 200 300 400 500 600 700 800
Pressure (atm)
60
700 atm
(b) 400 atm
200 atm
100 atm
50 50 atm
25 atm
Solubility (cc/grWater)

40

30

20

10

0
50 75 100
Temperature (C)

Fig. 8.22 Solubility of CO2 in water a isotherm solubility b isobar solubility [26]

increased. Also, they reported the relation between the concentration of H2 S in vapor
and liquid phase at different temperatures which can be seen in Fig. 8.24 [30].
Selleck et al. (1952) studied the phase behavior of the H2 S-water system. They
also reported the experimental data of the solubility of the H2 S in water which is
reported in Table 8.3 of their paper [31].
384 A. Izadpanahi and R. Azin

900
(a)
800

700
12 C
600
Pressure (atm)

40 C

500 75 C

100 C
400

300

200

100

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
Mole Fraction

40
(b)
35
Solubility (cc/grWater)

30

25

12 C
20
18 C
15 25 C
31.04 C
10
35 C
40 C
5

0
0 100 200 300 400 500 600
Pressure (atm)

Fig. 8.23 Solubility of CO2 in water a mole fraction of CO2 in water b solubility isotherms [27]
8 Water-Hydrocarbons System 385

Table 8.6 Solubility of CO2 in water phase reported by Chapoy et al. [29]
T(K) P (psi) CO2 mole fraction T(K) P (psi) CO2 mole fraction
274.14 27.5572 0.00262 298.4 403.204806 0.0142
274.83 174.19 0.0142 299.06 146.1980016 0.0059
276.74 192.465 0.0142 299.32 65.5570404 0.00262
278.22 33.0686 0.00262 303.99 72.3738123 0.00262
278.74 206.824 0.0142 310.86 1060.080549 0.02017
283.38 251.205 0.0142 311.62 91.373751 0.00262
284.27 41.6258 0.00262 313.36 597.4102863 0.0142
284.73 571.158 0.02488 321.64 112.9843683 0.00262
288.41 47.7174 0.00262 321.97 756.5166432 0.0142
289.2 299.068 0.0142 322.14 1353.636854 0.02017
289.62 702.563 0.02488 330.6 132.4194201 0.00262
292.35 750.135 0.02488 341.24 157.5109422 0.00262
293.01 340.694 0.0142 351.31 180.2818611 0.00262
293.34 55.8395 0.00262

0.2

0.18

0.16
Concentration in Vapor

0.14

0.12

0.1

0.08
5C
0.06
20 C
0.04 30 C

0.02 50 C

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Concentration in Solution

Fig. 8.24 H2 S concentration in vapor phase versus liquid phase [30]


386 A. Izadpanahi and R. Azin

Suleimanov and Krupp (1994) reported experimental solubility data of the H2 S


in pure water from 20 to 320 °C, at saturation pressure. They stated that Henry’s
constant has good agreement with the previously published data at the temperatures
below 200 °C. But, it is obtained higher solubilities at higher temperatures. They
reported the experimental solubility data of H2 S in pure water in their paper [32].
Chapoy et al. (2005) also reported solubility data of H2 S in water between 298.16
and 338.34 K and pressures up to 575 psi. Their experimental data are based on the
static analytic apparatus which taking advantage of two ROLSI pneumatic capillary
samplers. Also, they reviewed the reported H2 S solubility data in the literature. They
also proposed a thermodynamic model to represent the experimental data. Table 8.7
shows the solubility data reported by Chapoy et al. They compared the results of
their work with the literature data in Figs. 8.4 and 8.5 of their work [33].
Goodman and Krase (1931) developed a procedure for the determination of gas
solubilities at high pressures and temperatures. They reported the nitrogen solubilities
at pressures between 100 and 300 atmospheres and temperatures between 0 and
170 °C. They also calculated the coefficient in Henry’s law for several pressures.
Table 1 of their paper shows the solubility data for Nitrogen reported by Goodman
and Krase [34].
Wiebe et al. (1933) reported experimental data of the Nitrogen solubility in water.
The range of pressure in their work is wider (25–1000 atm) than Goodman and
Krase [34]. Figure 8.25a shows the isobar solubility of Nitrogen in water. As can

Table 8.7 Solubility of H2 S in water phase reported by Chapoy et al. [33]


T(K) P (psi) H2 S mole fraction T(K) P (psi) H2 S mole fraction
×103 ×103
298.16 72.954 9.12 328.28 72.0837369 4.96
298.16 100.076 12.78 328.28 146.1980016 10.05
298.16 115.595 14.69 328.28 217.2664746 14.89
308.2 70.0532 6.69 328.28 286.8845706 19.5
308.2 110.664 10.69 328.28 357.9530436 24.41
308.2 173.03 17.09 328.28 440.0443818 29.5
308.2 253.526 24.73 328.28 504.0060075 33.56
308.2 315.457 30.36 338.34 73.8241893 4.35
308.2 360.129 34.01 338.34 77.7402072 4.59
318.21 73.5341 5.81 338.34 152.7246981 9.23
318.21 152.725 12.5 338.34 244.8236376 14.71
318.21 218.572 17.87 338.34 321.2585055 19.15
318.21 293.556 23.9 338.34 405.5254092 24.2
318.21 310.236 24.72 338.34 488.777049 28.82
318.21 372.747 29.59 338.34 574.6393674 31.83
318.21 448.747 35.07
8 Water-Hydrocarbons System 387

8
(a)
7
1000 atm
800 atm
Solubility (cc/grWater)

6 500 atm
200 atm
50 atm
5 25 atm

0
0 25 50 75 100
Temperature (C)
8
(b)
7
25 C
Solubility (cc/grWater)

6 50 C
75 C
5 100 C

0
0 100 200 300 400 500 600 700 800 900 1000
Pressure (atm)

Fig. 8.25 Solubility of N2 in water a isobar solubility b isotherm solubility [35]

be seen, the solubility of nitrogen first decreases and then increases by increasing
the temperature. So, the data show a minimum at about 70 °C. This trend is more
obvious in Fig. 8.25b which shows the isotherm solubility of nitrogen in water. As it
can be seen, the solubility of Nitrogen in water at 100 °C is higher than the solubility
388 A. Izadpanahi and R. Azin

at 50 °C and lower than the solubility in 25 °C. Also, can be understood that for all
temperatures the solubility increases by increasing the pressure [35].
Saddington and Krase (1934) reported the solubility of Nitrogen in water. Their
results of experiments are shown in Fig. 8.26. Figure 8.26a shows the isobar solubility
of Nitrogen in water. The minimum solubility can be observed in the same range of
temperature which reported by Wiebe et al. [35]. Also it can be seen that in high
temperatures, the solubility of Nitrogen increases linearly by increasing temperature.
Figure 8.26b shows the isotherm solubility of Nitrogen in water which increases by
increasing pressure [36].
Chapoy et al. (2004) reported experimental data of Nitrogen in water. The exper-
imental data were compared with the available solubility data in the literature which
have good agreement. They used their experimental data to tune the binary interac-
tions between Nitrogen and water. They reported the experimental data of nitrogen
solubility in the water in Table 3 of their paper [37].

8.3.3 Calculation of Gas Solubility

Henry’s law is one of the calculation methods of gas solubility when the concentration
of components in the mixture is low. Equation (8.5) shows the proportionality of
component fugacity to its concentration [38].

f i = Hi xi (8.5)

where f i , H i and x i are the fugacity, henry’s constant and mole fraction in liquid phase
of component i, respectively. The assumption of ideal gas is valid at low pressure. In
these conditions, Eq. (8.6) can be written as follows [38]:

P y i = Hi xi (8.6)

Equation (8.6) is known as henry’s law. In this equation, P and yi are pressure and
the mole fraction of component i in the gas phase, respectively [38].
Henry’s constant of a component is experimentally determined. Also, it is assumed
to be independent of its concentration but, this constant is a function of temperature.
Figure 8.27 shows the Henry’s constant for gaseous components of reservoir fluids
in water [38].
Besides, the Henry’s constant dependency on pressure can be calculated using the
following relation:


Hi = Hi 0 exp vi ∞ P − P 0 /RT (8.7)
8 Water-Hydrocarbons System 389

7
(a) 300 atm
6
200 atm
Solubility (cc/grWater)

5 100 atm

0
0 50 100 150 200 250

Temperature (F)
7
(b)
6
230 C
Solubility (cc/grWater)

200 C
5 150 C
50 C
4 80 C

0
0 50 100 150 200 250 300 350

Pressure (atm)

Fig. 8.26 Solubility of N2 in water a isobar solubility b isotherm solubility [36]


390 A. Izadpanahi and R. Azin

13000
(a)
12000 Nitrogen
Propane
Henry's Constant (MPa/Mole Fraction)

11000 Hydrogen
Ethane
10000 Methane

9000

8000

7000

6000

5000

4000
280 300 320 340 360 380 400 420 440

Temperature (K)
600
(b)
Carbon Dioxide
500 Hydrogen Sulfide
Henry's Constant (MPa/Mole Fraction)

400

300

200

100

0
280 300 320 340 360 380 400 420 440
Temperature (K)

Fig. 8.27 Henry’s constant for different gases [39]

where Hi 0 and vi ∞ are Henry’s constant at P 0 and the partial molar volume of
component i in the solvent at infinite dilution. vi ∞ is assumed to be constant over
the prevailing composition and pressure ranges. Average of partial molar volume for
methane, ethane, propane and nitrogen are 40, 55, 80 and 35 cm3 /gmole, respectively
[38].
8 Water-Hydrocarbons System 391

Example 8.9 Estimate the methane solubility in water using Henry’s law in
the following conditions:
T = 350 K = 170 F and P = 41.36 MPa = 6000 psi.
Compare the results with experimental data using Fig. 8.18b.
Solution:
At T = 350 K, (using Fig. 8.27) H0 = 6400 MPa/mole fraction and P0 =
0.1 MPa. The Henry’s constant for methane is calculated using Eq. (8.7).

(41.36−0.1)
× exp 40 × 10−3 kgmmol ×
3
HC2 = 6400 moleMPa
fraction m3
=
(0.0083144×350) kg mol
11284.59 moleMPa
fraction
The solubility of ethane can be calculated using Eq. (8.6).
f C2 V = P y C2 φC2 V = HC2 xC2
The gas can be assumed as pure ethane because of the low volatility of water
in comparison to ethane: yC2 = 1. Also, it is assumed that φC2 V = 1. The gas
solubility is as follows:
xC2 = HPC2 = 11284.5941.36 MPa
MPa = 0.003665 Mole Fraction
mole fraction
Based on Fig. 8.18b, the experimental value of solubility is 0.0034.
In this example, the accuracy of results can be improved by calculating the
φC2 V using equation of state.

8.3.4 Effect of Salt on the Gas Solubility in Water

Connate water in reservoirs contains different concentration of different salts. The


presence of salt decreases the gas solubility in formation water. Several researchers
have studied the solubility of various gases in the brine. Soreide and Whitson provided
a simple approach for predicting the solubility of hydrocarbon gases and non-
hydrocarbon (N2 , CO2 , and H2 S) gases-brine system. In their approach, equation
of state (EOS) is used to predict the solubility at high pressure and temperature by
considering the effect of salinity. They validated their model using the experimental
data. The results showed that their proposed model could predict the experimental
data with high accuracy [40]. Smith et al. studied the solubility of nitrogen at pressures
below 1000 psia and room temperature. They measured solubility in the presence of
NaCl, CaCl2 , Na2 SO4 and MgSO4 in the solution. They concluded that the solubility
of nitrogen increases by increasing pressure. But, by increasing the salt concentration
the solubility of nitrogen decreases [41].
Duan et al. developed a model for the prediction of H2 S solubility in brine at
the pressure below 300 bar at 320 °C. They introduced an EOS for H2 S and H2 O
binary. This EOS can calculate chemical potentials of both H2 S and H2 O in vapor or
liquid phase. Below 350 °C NaCl has a significant effect on the activities of H2 S in
the liquid phase but, this effect is negligible on the vapor phase. Their model results
392 A. Izadpanahi and R. Azin

0.05
90 C
Henry's Constant 90 C
60 C
0.04 Henry's Constant 60 C

0.03
x H2S

0.02

0.01

0
0 10 20 30 40 50 60 70
Pressure (bar)

Fig. 8.28 Duan et al. results [42]

show good agreement with the experimental data. Figure 8.28 compares the results
of their model with the experimental data as well as the Henry’s constant calculation
[42].
Zirrahi et al. presented a new method for prediction of mutual solubility of CH4 ,
H2 S and CO2 and their mixtures in brine. They used Peng-Robinson EOS with a
non-random mixing rule for modelling acid and sour gas phase. The results showed
that their model can predict the solubility and water content at different conditions,
different kind of salt and different concentrations of salt. Figure 8.29a, b show the
solubility of H2 S in the 4 and 6 molarity of NaCl, respectively. As it is obvious from
this figure, the solubility decreases by increasing the salt concentration [43].

8.4 Activity Models

For defining the chemical potential of nonideal solutions or real solutions, a modifi-
cation has to be added in the expression of the chemical potential of an ideal solution.
This modification is added by using the activity coefficient which makes the chemical
potential expression as follows:

μi (T.P.x) = μi 0 (T.P) + RT ln γi (T.P.x)xi (8.8)

In this expression, γi is defined as the activity coefficient which is a function of


temperature, pressure and composition. In the case of x i → 1, the solution approaches
ideal behavior and all of the components in pure state are in the same state as the
8 Water-Hydrocarbons System 393

0.03
(a)

0.025

0.02
x H2S

0.015

0.01

313.15 K
333.15 K
0.005
353.15 K
393.15 K
0
0 2 4 6 8 10
Pressure (MPa)

0.025
(b)

0.02

0.015
x H2S

0.01

313.15 K
0.005 333.15 K
353.15 K
393.15 K
0
0 2 4 6 8 10
Pressure (MPa)

Fig. 8.29 The solubility of H2 S in brine a 4 molarity of NaCl b 6 molarity of NaCl [43]
394 A. Izadpanahi and R. Azin

mixture. In these conditions, the activity coefficient approaches to 1. Activity coeffi-


cient is a very useful function to define liquid phases in phase equilibrium calculation
[44].
There are several models which can be used to estimate the activity coefficient.
Prausnitz et al. extensively studied the activity models in their book. Also, they
reported the estimation of the activity coefficient at infinite dilution as well as the
group contribution models. More information of each model is available in the
Prausnitz et al. [45].
Zirrahi et al. [43] calculated the activity coefficient for aqueous CO2 , H2 S and CH4
using Duan and Sun [46], Duan et al. [47] and Duan and Mao [48]. They reported
that the Duan and Mao model for calculating the activity coefficient was used and
the predicted solubility of methane in water was very close to that the Duan and Mao
was predicted. Zirrahi et al. studied the water content prediction of sour and acid
gases. They stated in their work that the solubility of acid gases in water is low. Also,
they expressed that the activity of the water component can be approximated by its
mole fraction [49].

References

1. Campbell JM, Gas conditioning and processing, The basic principles, vol. 1, 9th ed. Norman,
Oklahoma: Campbell Petroleum Series; 2014.
2. G. P. S. A. (U.S.), G. P. Association. Engineering data book : SI version. Tulsa, Okla.: Gas
Processors Suppliers Association; 2012.
3. Moshfeghian M, Lean sweet natural gas water content correlation. PetroSkills TOTM; 2014.
https://www.petroskills.com/blog/entry/lean-sweet-natural-gas-water-content-correlation.
4. Moshfeghian M, Propane–Water phase behavior at low to moderate pressures. PetroSkills
TOTM; 2018. http://www.jmcampbell.com/tip-of-the-month/2018/09/propane-water-phase-
behavior-at-low-to-moderate-pressures/.
5. I. Bryan Research and Engineering, ProMax 3.2, Bryan, Texas; 2014.
6. Moshfeghian M. Ethane–Water phase behavior at low to moderate pressures. PetroSkills
TOTM; 2018. http://www.jmcampbell.com/tip-of-the-month/2018/12/ethane-water-phase-beh
avior-at-low-to-moderate-pressures/.
7. Moshfeghian M. Acid–gas water content. PetroSkills TOTM; 2014. http://www.jmcampbell.
com/tip-of-the-month/2014/02/acid-gas-water-content/.
8. Carroll JJ, The water content of acid gas and sour gas from 100 to 220 F and pressures to
10,000 Psia. In: 81st annual GPA convention; 2002.
9. Yarrison M. Water content of high pressure, high temperature methane, ethane and methane +
CO2 , ethane + CO2 Gas Processors Association; 2008.
10. Erbar JH, Jagota AK, Muthswamy S, Moshfeghian M. Predicting synthetic gas and natural
gas thermodynamic properties using a modified Soave-Redlich-Kwong equation of state. Gas
Process. Res. Report, GPA RR-42, Tulsa, USA; 1980.
11. Maddox RN, Lilly LL, Moshfeghian M, Elizondo E. Estimating water content of sour natural
gas mixtures. In: Laurance Reid gas conditioning conference, vol. 8; 1988.
12. Wichert GC, Wichert E. Chart estimates water content of sour natural gas. Oil Gas J (United
States). 1993;91(13).
13. Huang SS-S, Leu A-D, Ng H-J, Robinson DB. The phase behavior of two mixtures of methane,
carbon dioxide, hydrogen sulfide, and water. Fluid Phase Equilib. 1985;19(1–2):21–32.
8 Water-Hydrocarbons System 395

14. Moshfeghian M. Water-sour natural gas phase behavior. PetroSkills TOTM; 2007. http://www.
jmcampbell.com/tip-of-the-month/2007/11/water-sour-natural-gas-phase-behavior/.
15. Ng HJ, Chen CJ, Schroeder H. Water content of natural gas systems containing acid gas: GPA
Project 945. Gas Processors Association; 2001.
16. Culberson OL, Horn AB, McKetta JJ Jr. Phase equilibria in hydrocarbon-water systems. J Pet
Technol. 1950;2(01):1–6.
17. Culberson OL, McKetta JJ Jr. Phase equilibria in hydrocarbon-water systems II-The solubility
of ethane in water at pressures to 10,000 psi. J Pet Technol. 1950;2(11):319–22.
18. Culberson OL, McKetta JJ Jr. Phase equilibria in hydrocarbon-water systems III-the solubility
of methane in water at pressures to 10,000 psia. J Pet Technol. 1951;3(08):223–6.
19. Reamer HH, Sage BH, Lacey WN. Phase equilibria in hydrocarbon systems. n-butane-water
system in the two-phase region. Ind Eng Chem. 1952;44(3):609–15.
20. Song KY, Feneyrou G, Fleyfel F, Martin R, Lievois J, Kobayashi R. Solubility measurements of
methane and ethane in water at and near hydrate conditions. Fluid Phase Equilib. 1997;128(1–
2):249–59.
21. Dhima A, de Hemptinne J-C, Moracchini G. Solubility of light hydrocarbons and their mixtures
in pure water under high pressure. Fluid Phase Equilib. 1998;145(1):129–50.
22. Chapoy A, Mohammadi AH, Richon D, Tohidi B. Gas solubility measurement and modeling for
methane–water and methane–ethane–n-butane–water systems at low temperature conditions.
Fluid Phase Equilib. 2004a;220(1):113–21.
23. Chapoy A, Mokraoui S, Valtz A, Richon D, Mohammadi AH, Tohidi B. Solubility measurement
and modeling for the system propane–water from 277.62 to 368.16 K. Fluid Phase Equilib.
2004b;226:213–20.
24. Reshadi P, Nasrifar K, Moshfeghian M. Evaluating the phase equilibria of liquid water+ natural
gas mixtures using cubic equations of state with asymmetric mixing rules. Fluid Phase Equilib.
2011;302(1–2):179–89.
25. Mohebbi V, Naderifar A, Behbahani RM, Moshfeghian M. Determination of Henry’s law
constant of light hydrocarbon gases at low temperatures. J Chem Thermodyn. 2012;51:8–11.
26. Wiebe R, Gaddy VL. The solubility in water of carbon dioxide at 50, 75 and 100, at pressures
to 700 atmospheres. J Am Chem Soc. 1939;61(2):315–8.
27. Wiebe R, Gaddy VL. The solubility of carbon dioxide in water at various temperatures
from 12 to 40 and at pressures to 500 atmospheres. Critical phenomena. J Am Chem Soc.
1940;62(4):815–7.
28. Yang SO, Yang IM, Kim YS, Lee CS. Measurement and prediction of phase equilibria for
water+ CO2 in hydrate forming conditions. Fluid Phase Equilib. 2000;175(1–2):75–89.
29. Chapoy A, Mohammadi AH, Chareton A, Tohidi B, Richon D. Measurement and modeling of
gas solubility and literature review of the properties for the carbon dioxide−water system. Ind
Eng Chem Res. 2004;43(7):1794–802.
30. Wright RH, Maass O. The solubility of hydrogen sulphide in water from the vapor pressures
of the solutions. Can J Res. 1932;6(1):94–101.
31. Selleck FT, Carmichael LT, Sage BH. Phase behavior in the hydrogen sulfide-water system.
Ind Eng Chem. 1952;44(9):2219–26.
32. Suleimenov OM, Krupp RE. Solubility of hydrogen sulfide in pure water and in NaCl solutions,
from 20 to 320 C and at saturation pressures. Geochim Cosmochim Acta. 1994;58(11):2433–44.
33. Chapoy A, Mohammadi AH, Tohidi B, Valtz A, Richon D. Experimental measurement and
phase behavior modeling of hydrogen sulfide-water binary system. Ind Eng Chem Res.
2005;44(19):7567–74.
34. Goodman JB, Krase NW. Solubility of nitrogen in water at high pressures and temperatures.
Ind Eng Chem. 1931;23(4):401–4.
35. Wiebe R, Gaddy VL, Heins C Jr. The solubility of nitrogen in water at 50, 75 and 100 from 25
to 1000 atmospheres. J Am Chem Soc. 1933;55(3):947–53.
36. Saddington AW, Krase NW. Vapor—liquid equilibria in the system nitrogen—water. J Am
Chem Soc. 1934;56(2):353–61.
396 A. Izadpanahi and R. Azin

37. Chapoy A, Mohammadi AH, Tohidi B, Richon D. Gas solubility measurement and modeling for
the nitrogen+ water system from 274.18 K to 363.02 K. J Chem Eng Data. 2004;49(4):1110–5.
38. Danesh A. PVT and phase behaviour of petroleum reservoir fluids. Elsevier; 1998.
39. Kobayashi R, Katz D. Vapor-liquid equilibria for binary hydrocarbon-water systems. Ind Eng
Chem. 1953;45(2):440–6.
40. Søreide I, Whitson CH. Peng-Robinson predictions for hydrocarbons, CO2 , N2 , and H2 S with
pure water and NaCl brine. Fluid Phase Equilib. 1992;77:217–40.
41. O’Sullivan TD, Nagy B. Solubility of natural gases in aqueous salt solutions—III Nitrogen in
aqueous NaCl at high pressures. Geochim Cosmochim Acta. 1966;30(6):617–9.
42. Duan Z, Møller N, Weare JH. Prediction of the solubility of H2S in NaCl aqueous solution: an
equation of state approach. Chem Geol. 1996;130(1–2):15–20.
43. Zirrahi M, Azin R, Hassanzadeh H, Moshfeghian M. Mutual solubility of CH4 , CO2 , H2 S, and
their mixtures in brine under subsurface disposal conditions. Fluid Phase Equilib. 2012;324:80–
93.
44. Firoozabadi A. Thermodynamics of hydrocarbon reservoirs. McGraw-Hill; 1999.
45. de Azevedo EG, Lichtenthaler RN, Prausnitz JM. Molecular thermodynamics of luids-phase
equilibria; 1986.
46. Duan Z, Sun R. An improved model calculating CO2 solubility in pure water and aqueous NaCl
solutions from 273 to 533 K and from 0 to 2000 bar. Chem Geol. 2003;193(3–4):257–71.
47. Duan Z, Sun R, Liu R, Zhu C. Accurate thermodynamic model for the calculation of H2 S
solubility in pure water and brines. Energy Fuels. 2007;21(4):2056–65.
48. Duan Z, Mao S. A thermodynamic model for calculating methane solubility, density and gas
phase composition of methane-bearing aqueous fluids from 273 to 523 K and from 1 to 2000
bar. Geochim Cosmochim Acta. 2006;70(13):3369–86.
49. Zirrahi M, Azin R, Hassanzadeh H, Moshfeghian M. Prediction of water content of sour and
acid gases. Fluid Phase Equilib. 2010;299(2):171–9.
Chapter 9
Challenges of Gas Injection

Reza Azin, Amin Izadpanahi, and Ali Ranjbar

Abstract Gas injection operations are faced with important challenges. These chal-
lenges must be studied carefully before the operation in order to increase the gas
injection efficiency. This chapter provides a review of the most important issues for
designing the different gas injection methods. The compatibility of fluids and rocks
is discussed after the introduction part. In Sect. 9.3, corrosion in the different gas
injection methods such as carbon capture and sequestration (CCS), acid and flue gas
injection are investigated. Gravity override as one of the vigorous problems of gas
injection is discussed in Sect. 9.4. Gas mobility control procedures are introduced
in Sect. 9.5. In Sect. 9.6, cap rock integrity as one of the important issues in the gas
injection operation is studied in detail. Phase trapping during the gas injection and
HSE are discussed in Sects. 9.7 and 9.8, respectively.

Nomenclature

Acronyms

CCS Carbon capture and sequestration


EOR Enhanced oil recovery
FCM First contact miscibility
HSE Health, Safety and Environment

R. Azin (B) · A. Ranjbar


Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir
A. Ranjbar
e-mail: ali.ranjbar@pgu.ac.ir
A. Izadpanahi
Oil and Gas Research Center, Persian Gulf University, Bushehr, Iran
e-mail: a.izadpanahi@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 397
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_9
398 R. Azin et al.

MCM Multiple contact miscibility


MVA Monitoring, verification and accounting
SAG Surfactant-alternating gas
SRF Screening and ranking framework
SIMGAP Simultaneous injection of miscible gas and polymer
WAG Water alternating gas
UGS Underground Gas Injection

Variables

g Gravitational acceleration
M Mobility Ratio
NG Gravity Number
p Partial pressure, bar
T Temperature, K
Vcor Corrosion rate, mm/yr

Greek Letters

ρ Specific Gravity
μ Viscosity

Subscripts

o Displaced fluid
s Displacing fluid

9.1 Introduction

Gas injection operations are faced with important challenges. These challenges must
be studied carefully before the operation in order to increase the gas injection effi-
ciency. This chapter provides a review of the most important issues for designing the
different gas injection methods. The compatibility of fluids and rocks is discussed
after the introduction part. In Sect. 9.3, corrosion in the different gas injection methods
such as carbon capture and sequestration (CCS), acid and flue gas injection are inves-
tigated. Gravity override as one of the vigorous problems of gas injection is discussed
9 Challenges of Gas Injection 399

in Sect. 9.4. Gas mobility control procedures are introduced in Sect. 9.5. In Sect. 9.6,
cap rock integrity as one of the important issues in the gas injection operation is
studied in detail. Phase trapping during the gas injection and HSE are discussed in
Sects. 9.7 and 9.8, respectively.

9.2 Fluid-Fluid and Rock-Fluid Compatibility

One of the important subjects in designing gas injection projects is fluid-fluid and
rock-fluid compatibility. The incompatibility between them can cause the formation
damage which decreases the production rate or well injectivity and overall efficiency
of the gas injection project. Many researchers studied the formation damage during
CO2 injection in both EOR and storage processes [1–7]. CO2 injection as an enhanced
oil recovery (EOR) method can cause the precipitation of heavy components like
asphaltenes. Permeability and porosity are reduced due to asphaltene precipitation
[8, 9]. Also, CO2 reacts with the carbonate rocks which causes the calcite precipitation
and as a result permeability and porosity reduce. In carbon dioxide storage in saline
aquifers, the well injectivity is influenced by the reactions of CO2 with formation
rocks and formation water. The calcite precipitation which causes the fines migration
can reduce the permeability and porosity of the formation rocks [1]. During the acid
gas injection, rock permeability and porosity alteration happen due to the mechanical
and chemical mechanisms such as dissolution, desiccation and fine migration [10].

9.3 Corrosion

Corrosion causes many problems in oil and gas industry which attack the produc-
tion, transfer and process equipment. O2 , H2 S, CO2 , naphtenic acids and aromatic
carboxylic acids are some of the corrosive components which can affect all the casings
and pipelines [11]. Operations in the upstream industry which faced the corrosive
components including the CCS, acid gas disposal, flue gas disposal and some of
enhanced oil recovery methods.

9.3.1 CCS

CCS projects are encountered significant metal loss due to corrosion by CO2 . Water
phase behavior and the availability of an aqueous medium is an important parameter
in carbon dioxide corrosion. Studies show that carbon steel equipment can be used
for handling high CO2 partial pressures if the operation pressure is above the dew
point and the liquid water quantity is limited. Also, carbon dioxide can be caused the
corrosion rate of 10–20 mm/yr when water condensation happens. Other affecting
400 R. Azin et al.

parameters on CO2 corrosion include temperature, flow rate, alloy composition and
microstructure, solution pH, CO2 partial pressure, H2 S and O2 content. The detailed
effect of these parameters can be found in Chambers et al. [12].
The steps of identify and quantify corrosion damage in CCS systems are reported
by Chambers et al. as follow [12]:
Step 1. Phase behavior and streams thermodynamic modelling.
Step 2. Specifying the key parameters that affect the corrosion.
Step 3. Corrosion rates calculation for the worst case through prediction models
and/or simulated service environment testing.
Step 4. Defining the importance and associated risk of streams, equipment and
critical processes.
Step 5. Estimation of corrosion-resistance alloys base on the forecasted corrosion
rate.
The corrosion reactions in the case of CCS can be expressed as follow [12]:

Fe + 2H2 CO3 → Fe2+ + 2HCO−


3 + H2

Fe2+ + 2HCO−
3 → FeCO3 + H2 O + CO2

The calculation of corrosion rate for a specified range of conditions introduced


by DeWaard and Milliams as [13]:

1710
log(Vcor ) = 5.8 − + 0.67 log( pC O2 ) (9.1)
T
where V cor , T and pCO2 defined as corrosion rate (mm/yr), operating temperature
(°K) and partial pressure of CO2 (bar), respectively.

9.3.2 Acid Gas Disposal

A mixture of CO2 and H2 S is known as the acid gas which is a product of gas
sweetening process of sour gas. Dispose of these harmful gases is the purpose of
acid gas injection. There is no difference between the methods of disposing the acid
gas and CCS methods. The depleted oil and gas reservoirs and saline aquifers are
selected as a place for storing acid gas. Due to the nature of H2 S which is more
corrosive and toxic than CO2 , the preparation of large scale and high technologies
for CO2 geological storage are required for the success of acid gas disposal [14].
The injection of acid gas must have four circumstances to minimize risk and
optimize disposal include: (1) Acid gas must be injected into a dense fluid phase for
increasing the capacity, reducing the buoyancy and preventing leakage and migration,
(2) The injection pressure must be bigger than the formation pressure, (3) Preventing
9 Challenges of Gas Injection 401

the hydrate formation by increasing the system temperature above 35 °C in order to


mitigate the plugging of the pipeline and disposal well and (4) The water content
must be below the saturation limit in order to prevent corrosion [15, 16].
The possibility of corrosion and hydrate formation can be minimized by gathering
the water content and retaining the temperature above the hydrate stability temper-
ature. In this case, the transportation and injection of acid gas can be done using
the less expensive steels. Higher temperatures can be avoided hydrate formation and
plugging and also can be accelerated corrosion. Containment failure can be caused
by both corrosion and plugging. Chemical inhibitors like glycol and methanol and
corrosion resistant steels reduce the possibility of corrosion and plugging [17].

9.3.3 Flue Gas Injection

N2 , CO2 , O2 , H2 S, hydrocarbons and water vapor form the flue gas which used for
EOR. Metal and alloys corrosion may occur during the presence of CO2 , O2 , and H2 S
in a wet environment. So, flue gas injection is faced with the high risk of corrosion
in both surface and sub-surface equipment. In order to minimizing the corrosion
caused by flue gas injection the gas pressure must be kept above its dew point and
the Oxygen contents must be minimized [18].
Two kinds of corrosion exist including sweet and sour corrosion which caused by
CO2 and H2 S, respectively. CO2 corrosion is more critical than HCl corrosion due
to the higher acidity when it dissolves in the water. The existence of SO2 and O2
in the flue gas will speed up the corrosion. O2 comes from the excessive air of the
boiler and SO2 is generated from the combustion of oil and coal [19, 20]. Some of
the corrosion products of flue gas injection include FeCO3 , FeO(OH), FeS, Fe3 O4
and S. More information about the reactions and the corrosion products of flue gas
injection could be found in the Zhong et al. [21].
Li et al. comprehensively studied the gas and liquid flue gas corrosion of four
materials such as X70, P110, N80 and 13Cr. The downhole strings consist of these
materials. They reported the influencing factor on the corrosion rate of flue gas injec-
tion include corrosion time, temperature, pressure, velocity, O2 , SO2 , H2 O and NaCl
concentrations. In the gas phase, corrosion rate of N80, P110 and X70 increase signif-
icantly with increasing the temperature and O2 concentration. Firstly, the corrosion
rate increases by increasing the pressure. But, after a specific amount of pressure, the
corrosion rate becomes constant. In the liquid phase, the corrosion rates of N80, P110
and X70 increase and then reduce with increasing the temperature. Also, increasing
the velocity and the concentrations of SO2 , NaCl, SO2 and O2 , significantly increase
the corrosion rates of N80, P110 and X70. 13Cr has shown good corrosion resis-
tance performance in comparison to other materials due to lower corrosion rate. They
suggested that low humidity of flue gas, low concentration of O2 , avoid co-injection
of flue gas and brine and the temperature lower than 80 and higher than 100 will
decrease the equipment corrosion [19].
402 R. Azin et al.

9.4 Gravity Override

One of the vigorous problems in gas injection processes is gravity override. This
phenomenon will happen when the density of injection gas is lower than the
displacing fluid and the vertical communication is high. Gravity segregation occurs
in this situation and gas sweeps the top of the reservoir. Gravity override is not a
challenging problem in stratified and in the reservoirs with low vertical permeability
[22–26]. Gravity override decreases the recovery of first contact miscible (FCM) gas
flooding which causes the recovery of FCM much lower than vertical floods. The
absence of gravity override in the multiple contact miscible (MCM) gas flood causes
a higher recovery in comparison to FCM and slightly lower recovery in comparison
to vertical floods [27, 28].
Dimensionless density difference and gravity number are two numbers which
used to investigate the effect of gravity on miscible gas injection. These numbers are
as follows:
ρo − ρs
ρ = (9.2)
ρs
(ρo − ρs )gk
NG = (9.3)
qμo

where ρ o , ρ s , g, k, q and μo are the density of displaced and displacing fluids,


gravitational acceleration, permeability, flow rate and displaced fluid viscosity. The
effect of gravity becomes insignificant when the gravity number (Ng ) is small. By
increasing the gravity number, the growth rates of individual fingers are affected by
gravity as the viscous fingering happens. As the gas goes upward due to its lower
density, the fingers in the upper part of the reservoir grow faster. In contrast, the
growth of the lower fingers is stopped due to the gravity drainage and also, no more
gas enters these fingers. The results of this phenomenon are the early breakthrough
of gas and reduction in sweep efficiency.
A gravity tongue is grown on the top of the reservoir when the gravity number
is large enough. Near this tongue, viscous fingering may happen, while fingering in
other parts of the porous media is stopped. In this situation, both viscous fingering
and gravity override have significant impact on the displacement. Besides, the gravity
override completely overcomes the displacement when the gravity number is very
high and also the fingers that form in early times is suppressed. The rate of oil
recovery and the breakthrough time are very low in this case. It should be pointed
out that by increasing gravity number the breakthrough time and the recovery after
breakthrough are significantly decreased [29, 30].
CO2 injection in saline aquifers and also in enhanced oil recovery are faced gravity
override due to lower density and viscosity in comparison to formation fluids [31–34].
This gravity override in enhanced oil recovery gas flood decreases the sweep effi-
ciency and accelerates the breakthrough of injected CO2 . In this case, large amounts
of reservoir oil being unswept. Water-alternating gas (WAG) is one of the methods
9 Challenges of Gas Injection 403

which could overcome the gravity override. However, the segregation of gas and
water after injection is a severe problem in this method. Surfactant-alternating gas
(SAG) and foam-assisted WAG can increase the injectivity and the swept area of
the reservoir. Also, the horizontal producer method can use the advantage of density
difference between the injection gas and reservoir oil. In this method, the gas is
injected into the upper part of a reservoir through vertical wells and oil is produced
from the lower part of the reservoir through a horizontal well [34].

9.5 Mobility Control

Mobility is defined as the ratio of effective permeability to the viscosity of a fluid.


Mobility ratio (M) is one of the important factors that affect the displacement process
and is defined as the mobilities ratio of displacing to displaced fluid. Due to variation
of fluids viscosity, M is not constant during the displacement process. A major
problem of EOR gas flood is unfavorable mobility ratio which has to be controlled
for increasing the efficiency of the process. Unfavorable mobility ratio can cause the
fingering which is the major reason of gas flooding process failure. The problems
associated with mobility is being very complicated due to the heterogeneity of porous
media and the three phase flow instead of two phase. Water alternating gas (WAG),
gas-soluble viscofier, surfactants, foams and polymers injection are some methods
for controlling the unfavorable mobility [30]. These methods and other methods of
mobility control are described briefly in the following paragraphs.
WAG is known as a mobility control process for gas injection projects. However,
this method has faced some challenge include water blocking, mobility control in
heavy oil reservoirs, reduction of gas injectivity, gravity segregation and reduc-
tion of oil relative permeability. Chemically enhanced WAG has been proposed for
improving the efficiency of WAG process. In this method, a chemical slug consists of
surfactant, alkaline and polymer is injected after one cycle of gas and water slug. This
method can effectively increase the efficiency of WAG process due to decreasing oil
viscosity, interfacial tension (IFT) and water blocking and also, improving mobility
ratio [35].
Foam creation using surfactant alternating gas (SAG) is another method for
controlling the mobility ratio in a gas injection process [36–39]. High effectiveness,
low cost, unaffected by contact with crude oil and reservoir minerals and chemi-
cally stable, soluble and surface active in oilfield brines are the characteristics of
a useful surfactant for using as a mobility control agent in gas injection process.
Nonylphenol ethoxylate sulphonate (NPES) is one of the surfactants which used
for mobility control in CO2 injection. This surfactant effectively decreases the IFT
between CO2 and brines and also, decrease the mobility of CO2 [40, 41]. Alpha olefin
sulphonate as main and octylphenol ethylene oxide and lauryl amido propyl amine
oxide as additives in three formulations are also used for CO2 WAG. These agents
increase the pressure differential and as a result decrease the gas mobility [42].
404 R. Azin et al.

Shehadeh et al. comprise the in-situ foam generation with simultaneous injection
of miscible gas and polymer (SIMGAP). In the in-situ foam generation method, gas
and surfactant are injected in the lower zone and upper zone, respectively. Foam will
form in-situ as the gas migrates to the upper zone which causes the reduction of
gas mobility. Also, this method can retain the gas in the lower zone and improve the
vertical sweep efficiency. In SIMGAP method, polymer solution and miscible gas are
injected in the upper zone and the lower zone, respectively. Also, injected polymer
can cause a lateral pressure gradient in upper zone, helping the gas confinement
in lower zone. Their results show that SIMGAP method leads more recovery than
in-situ foam generation [43].

9.6 Cap Rock Integrity

In general, any factor that causes the injection gas (or hydrocarbon in the reservoir)
to escape or leak from the injection path, reservoir and cap rock will be the most
important hazard in the injection process. In fact, injected gas or hydrocarbon leakage
can occur from existing pathways (such as fractures in the cap rock) or induction
pathways due to injection (such as reactivation of the fault and expansion of fractures
in the cap rock or dissolution paths in it) [44]. There are two major perspectives on
the risks associated with CO2 injection (the goal of increasing recovery and storage).
In the first category, these risks are considered as time-dependent effects and it
is divided into short and long term effects. In the second category, their types or
mechanisms are considered which is divided into chemical and mechanical hazards
[45, 46]. Figure 9.1 shows the cap rock integrity problems which may occur due to
CO2 injection.

Fig. 9.1 Cap rock problems to CO2 injection


9 Challenges of Gas Injection 405

Production of hydrocarbons or injection of fluids into reservoir may cause the


stress change in cap rock and reservoir. New fractures will be formed or pre-existing
fractures and faults will be reactivated due to high stress changes. These fractures
may be threaten cap rock integrity [47]. One of the important issues in CO2 injection
is the absence of cap rock integrity which is known as cap rock failure. This issue
can cause the leakage of injected CO2 which in case all the operations, money and
time will be useless. Cap rocks define as the low permeable geological traps which
prevent the migration of fluids into upper formation. Cap rock failure may occur
due to some mechanisms including (1) capillary leakage (2) gas diffusion (3) man-
made (hydraulic) fractures (4) natural deformation and fracturing in the cap rock (5)
formation of high permeable zones due to chemical reactions (6) open faults and
(7) seismic disturbances. Fractures and faults are considered as the most important
mechanisms of cap rock failure [46, 48, 49].
As mentioned above, assessment of the risk of fracturing is a crucial step for plan-
ning a CO2 storage process. In-situ stresses, the existence of natural fracture network
and material strength and stiffness are the three main parameters that control the
fracturing. This type of failure can be characterized by coupled geo-mechanical and
reservoir modelling of the reservoir cap rock system [50, 51]. Specific investigation
on a CCS pilot in Norway shows that stiffness and strength of virginal material are
very high in the reservoirs with shallow depth. Also, fracturing is not very common
when the existence of natural fractures and high tensile strength of intact formation
are combined [52]. Temperature difference between the reservoir and injected CO2
can cause the creation or re-opening of the fractures. Compressive and tensile stresses
will be decreased by cooling the reservoir and cap rock with injected CO2 due to
inability of the rock to contract freely [53].
Injection pressure is an important operational parameter in maintaining the cap
rock integrity. High injection pressure can cause the damaging of cap rock and
destroy the sealing performance. Fracture-closure pressure is selected as the basis
for selecting the injection pressure. As it is known, induced fracture cannot be formed
in the pressures lower than the fracture-closure pressure. Also, the opening of existing
fractures cannot be increased in this situation. So, the maximum injection pressure
should be lower than fracture-closure pressure [54].
Several researchers studied the changes in reservoir and cap rock porosity and
permeability due to the CO2 storage. In the Ketzin pilot, the integrity of reservoir, cap
rock and fault are stable and the variation of porosity and permeability is negligible.
But, in the Nagaoka the porosity does not change and the permeability could increase
significantly by the changing of pore structure due to supercritical CO2 injection.
Also, it is reported that a small reduction in the porosity of the cap rock can improve
the sealing capacity. In contrast, the leakage of CO2 from the geological formation
can be increased by increasing the porosity [55–57].
A detailed study was conducted on the effects chemical reactions on the trapping
mechanisms, cap rock and well integrity for the Krechba CO2 storage site. Pore
space of the reservoir and cap rock may alter due to Al/Fe silicate dissolution. Also,
silicate dissolution and amorphous silica, smectite and boehmite precipitation will
be increased by decreasing the pH. Long term mineral trapping may be caused by
406 R. Azin et al.

the dissolved Fe. Besides, reservoir and cap rock permeability may be decreased due
to precipitation of clays, secondary Fe-carbonates and hydroxides [58].
In a particular case of acid gas disposal in Zama oil field, the cap rock failure would
not happen in the normal operation conditions. However, investigations indicate that
fluid flow process, mineralizing reactions and dissolution in overlying formation
waters would prevent the migration of acid gas into the surface if cap rock failure
occurs [59, 60].

9.7 Trapping

Phase trapping is defined as the reduction in oil and gas permeability and relative
permeability due to an increase in an existing phase saturation or introducing an
additional immiscible phase into the porous media [61]. It is essential to correctly
estimate the phase trapping during the gas injection processes. Incorrect estimation
of trapped fluids can lead to inaccurate timing, over- or under-estimation of CO2
required injection volume and also its amount stored in the reservoir and inaccurate
amount of oil recovery in EOR processes. In water-oil-gas systems, each of these
three phases can be trapped which is related to microscopic effects such as pore
doublet or snap-off trapping [62].
It is mentioned in the literature that water injection before gas injection can
increase the oil trapping. It is suggested that the gas flooding in a virgin oil reservoir
has higher recovery than gas flooding in a water-flooded reservoir. Also, for CCS
purpose gas injection into a virgin oil reservoir is more suitable than a water-flooded
one [63].
The experiments on vertical sandstone cores demonstrate that higher trapped gas
saturation will be existed in the porous layer compared to the dense layer at the end
of the gas injection. Also, the residual gas saturation in the dense layer is as same
as the gas saturation at the end of gas injection. In the case of horizontal core, the
gas saturation is lower than the vertical core due to the movements of injected gas
through permeable layers [64].
Capillary trapping is one of the favorable trapping mechanisms in CCS processes
which has high storage security in comparison to hydrodynamic, dissolution and
mineral trapping. In this mechanism, the gas is trapped in the geological formation
due to capillary forces. The investigations of CO2 injection in brine show that the
trapped gas saturation is increased as increasing the viscosity and density difference
between fluids. These results can be used for optimal trapping of CO2 by alteration
of viscosity or density [65]. Further information about trapping are presented in
Chap. 10.
9 Challenges of Gas Injection 407

9.8 Health, Safety and Environment (HSE) of Gas Injection

9.8.1 CCS

Increasing the concentration of CO2 in the air can cause some issues. These issues
may be short or long term, global or local and related to health, safety and envi-
ronment. Threatening life is the short term consequence of increasing the CO2
concentration. High concentration of CO2 in the air can increase the blood acidity
and has negative effects on central nervous system, cardiovascular and respiratory.
However, atmospheric concentration of CO2 is crucial for respiratory and photo-
synthetic processes. Ecosystem and environmental effects are categorized as long
term outcomes of CO2 leakage. CO2 may react with the rock, dissolve in brine and
remobilize problematic materials. Also, CO2 could alter the pH of water by forming
a weak carbonic acid. This alteration can induce geochemical reaction which can
mobilize the metal contaminants such as Pb and Hg. Marine life such as meso- and
bathy-pelagic species and zooplankton also be affected by the PH changes in the sea
by leaking CO2 into the seafloor. The risk of mentioned leakage is very low for any
individual operation. It should be noted that, by increasing the number of storage
projects the overall risks will be increased [66].
There are some international frameworks for CCS operations which studied by
Mace et al. The 1972 London convention, the 1996 London protocol, 1992 OSPAR
convention, UN framework convention on climate change, Kyoto protocol and EU
frameworks are the most important frameworks on CCS operations. However, some
of these frameworks will be limited the CCS operations. More information include the
details of the mentioned frameworks and existing gaps in large scale CCS operation
can be found in Mace et al. [67].
It has to be mentioned that, for operation of a CCS project in large scales, a full
field study must be conducted in order to specify the optimal scenarios of storage
and to mitigate the chance of project failure. Many researchers have worked on the
methods for screening and ranking framework (SRF) based on HSE risks [68, 69].
These methods can evaluate the risk of leakage on a CO2 storage site which will help
the operator to perform a successful project by reducing the risks.
Li et al. used the Oldenburg method [69] for evaluating the leakage risk of Shenhua
CCS project in Ordos basin, China. They revised the SRF method for assessing the
Shenhua CCS project and studied two scenarios. Based on their study, this basin can
be divided into six reservoir seal combinations with fair attenuation potential and
primary and secondary containment. Their results showed that this site has a low
leakage risk and it is suitable for CCS operation [70]. The short term risk assessment
of CCS in Shenhua and the risk management in this project are discussed by Diao et al.
[71]. Also, a detailed review of the HSE regulatory framework and environmental
risk management of CCS in China are available in the Seligsohn et al. and Liu et al.,
respectively [72, 73].
Lee and Choi studied the damage costs of CO2 leakage in a CCS project in
South Korea. They divided the damages of CO2 leakage into three categories include
408 R. Azin et al.

economic, environmental and human health. Economic damage consists of the


destruction of marine ecosystem and the required costs for the recovery facilities and
marine ecosystem. Harm to marine ecosystem and climate change due to increasing
the CO2 concentration can be formed the environmental damage. CO2 high concen-
tration can cause some disease such as, headache, dyspnea, dizziness and even lead
to die. They suggested that to support a more positive public attitude and to ease
public concerns, the safety of a CCS project should be monitored carefully by the
government [74].
Safety of CCS projects must be considered in all three stages of this operation
include capture, transport and injection. The capture stage hazards can be catego-
rized as fire or explosion due to loss of containment of Oxygen, fire due to loss
of containment of flammable solvents like amine which used for CO2 capture and
explosion as a result of loss of containment of synthesis gas. The low temperature
and high concentration toxicity of CO2 can be caused damage to people health in
the transport stage of a CCS project if the loss of containment occurs. Also, in the
injection stage the loss of containment can lead to health damage to people and fire
or explosion of hydrocarbon systems or platform structure [75].

9.8.2 EOR

HSE consideration must take into account for the EOR processes. CO2 injection is
one the most applicable EOR methods which the HSE issues related to them were
discussed in the previous section. Also, it is reported that reservoir, near surface,
production and injection should be monitored for mitigation the CO2 leakage in
EOR projects. Using corrosion inhibitors, tube fault detection, production dynamic
analysis and using anti-CO2 corrosion cement are suitable operations for the safety
at the well site [76].
The technical approach of Zama acid gas injection EOR project can be categorized
into two groups include monitoring, verification and accounting (MVA) program and
injection program. These two groups are related to each other and some activities and
data sets are common in both groups. MVA has three purposes include (1) demon-
strate the security of the operation by considering the relevant data sets about safety
and containment (2) investigation of the project effect during and after injection by
providing a set of baseline conditions and (3) preparing a technical framework to
increase the financial outcomes of the decreasing the emissions at Zama field. The
introduction of a tracer and data collection such as reservoir dynamic monitoring
and reservoir fluid sampling can be categorized as the field-based elements. Base-
line conditions, which mentioned before, are defined based on the geological and
hydrogeological characterization, reservoir characteristics, geochemical and geome-
chanical properties of the reservoir and cap rock system and evaluation of wellbore
integrity. Also, early identification of leakage can be achieved through pressure data
(historical and new pressure of the reservoir).The injection program purposes are
(1) financial aspect of the injection operations, (2) simplify the oil production and
9 Challenges of Gas Injection 409

(3) effective CO2 storage. The key aspects in the injection program include acid
gas injection and EOR operations, well preparation and maintenance activities and
infrastructure and capture elements of the project [77].

9.8.3 Underground Gas Injection (UGS)

In the underground gas storage operations, safety, gas tightness and monitoring of
gas storage foundation are the most important part of the operation. The escaped gas
from the underground can harm the people health. People fatality are reported in some
UGC operations. Impermeable rock formation at the top of target injection geological
layer which can play the role as a seal is a suitable situation for underground gas
storage operation. Impermeable formation is very beneficial because they improve
the bond between the casing and the cement due to creep naturally and tend to tighten
around the well. The situation for UGS may not work in depleted oil and gas field
with many drilling wells because the drilled wells completion may damage. These
conditions increase the chance of gas leakage through the drilled wells. Operational
pressure range, safety margin and abandonment procedure are some of the safety
consideration during the UGS operation in the caverns. Salt domes or halite beds can
lead to some problems in the presence of water. The most common problem is the
corrosion of well casings and cement. Mechanical integrity test is an important pre-
stage of storage in salt cavern. In this test, the pressure in the cavern is increased and
then, the wells are closed. The pressure variation in the cavern is monitored carefully
over time. A pressure cycling is needed for testing the cavern conditions for storage
and test the damage of cavern wall. The maximum pressure can be chosen by rock
mechanical test, temperature and depth of the cavern. It is reported in the literature
that salt cavern is the safest form of hydrocarbon storage and the risks of UGS can
be minimized if proper instructions are followed [78].
The cap-rock of the depleted oil and gas fields, which used as a gas storage
place, commonly saturated with water. Capillary water trapping is the dominated
mechanism of sealing the cap rock. Capillary threshold pressure is a parameter that
the below of this pressure the cap rock prevents the gas migration to the upper
formation. But in the pressures above the capillary threshold pressure, the cap rock
becomes permeable to gas due to the drainage of water [79, 80].
The injection of natural gas containing high concentration of Hydrogen can cause
some problems. Abiotic reactions, acidification and increasing microbial activities
are some of the issues that will be activated due to existence of Hydrogen. More
information can be found in the Reitenbach et al. [81].
Integrity issues such as well integrity and fault reactivation are the important
issues in the UGS operations. Completion techniques are different for the existing
and new drilled wells. Mechanical and chemical effects like corrosion of down-hole
equipment must be considered in the drilling and completion of new wells. Open-
hole gravel pack could improve the performance of existing wells. Lost circulation
control, tubing selection, retrievable downhole packer, subsurface safety valves and
410 R. Azin et al.

formation damages prevention are some of the technologies that used in the UGS
operations. As mentioned, the leakage pathways can be caused by integrity loss
which may significantly decrease the efficiency of the project [82].

References

1. Okwen RT. Formation damage by CO2 asphaltene precipitation. In: SPE International
symposium and exhibition on formation damage control. Society of Petroleum Engineers;
2006.
2. Jin M, Ribeiro A, Mackay E, Guimarães L, Bagudu U. Geochemical modelling of formation
damage risk during CO2 injection in saline aquifers. J Nat Gas Sci Eng. 2016;35:703–19.
3. Kalantari-Dahaghi A, Gholami V, Moghadasi J, Abdi R. Formation damage through asphaltene
precipitation resulting from CO2 gas injection in Iranian carbonate reservoirs. SPE Prod Oper.
2008;23:210–4.
4. Kalantari-Dahaghi AM, Moghadasi J, Gholami V, Abdi R. Formation damage due to asphaltene
precipitation resulting from CO2 gas injection in Iranian carbonate reservoirs. In: SPE Europe
annual conference and exhibition. Society of Petroleum Engineers; 2006.
5. Mohamed IM, Nasr-El-Din HA. Formation damage due to CO2 sequestration in deep saline
carbonate aquifers. In: SPE International symposium and exhibition on formation damage
control. Society of Petroleum Engineers; 2012.
6. Mohamed I, Nasr-El-Din HA. Fluid/rock interactions during CO2 sequestration in deep saline
carbonate aquifers: laboratory and modeling studies. SPE J. 2013;18:468–85.
7. Mohamed IM, He J, Nasr-El-Din HA. Permeability change during CO2 injection in carbonate
aquifers: experimental study. In: SPE Americas E&P health, safety, security, and environmental
conference. Society of Petroleum Engineers; 2011.
8. Srivastava RK, Huang SS, Dong M. Asphaltene deposition during CO2 flooding. SPE Prod
Facil. 1999;14:235–45.
9. Minssieux L, Nabzar L, Chauveteau G, Longeron D, Bensalem R. Permeability damage due to
asphaltene deposition: experimental and modeling aspects. Rev l’Institut Français Du Pétrole.
1998;53:313–27.
10. Bennion DB, Thomas FB, Ma T. Formation damage processes reducing productivity of low
permeability gas reservoirs. In: SPE Rocky Mountain regional/low-permeability reservoirs
symposium and exhibition. Society of Petroleum Engineers; 2000.
11. Tiu BDB, Advincula RC. Polymeric corrosion inhibitors for the oil and gas industry: design
principles and mechanism. React Funct Polym. 2015;95:25–45.
12. Chambers B, Kane RD, Yunovich M. Corrosion and selection of alloys for CCS (Carbon
Capture and Storage) systems: current challenges. In: SPE International conference on CO2
capture, storage, utility. Society of Petroleum Engineers; 2010.
13. De Waard C, Milliams DE. Carbonic acid corrosion of steel. Corrosion. 1975;31:177–81.
14. Bachu S, Gunter WD. Acid-gas injection in the Alberta basin, Canada: a CO2 -storage
experience. Geol Soc Lond Spec Publ. 2004;233:225–34.
15. Carroll JJ, Lui DW. Density, phase behavior keys to acid gas injection. Oil Gas J. 1997;95.
16. Ng H-J, Carroll JJ, Maddocks J. Impact of thermophysical properties research on acid gas injec-
tion process design. In: Proceedings of the annual convention of the gas processors association.
Gas Processors Association; 1999. p. 114–20.
17. Bachu S, Haug K, Michael K, Buschkuehle BE, Adams JJ. Deep injection of acid gas in Western
Canada. Dev Water Sci. 2005;52:623–35.
18. Yee C-T, Stroich A. Flue gas injection into a mature SAGD steam chamber at the Dover Project
(Formerly UTF). J Can Pet Technol. 2004;43.
19. Li S, Zhang K, Wang Q. Experimental study on the corrosion of a downhole string under flue
gas injection conditions. Energy Sci Eng. 2019;7:2620–32.
9 Challenges of Gas Injection 411

20. Nesic S. Effects of multiphase flow on internal CO2 corrosion of mild steel pipelines. Energy
Fuels. 2012;26:4098–111.
21. Zhong X, Wang Y, Liang J, Chen L, Song X. The coupling effect of O2 and H2 S on the corrosion
of G20 steel in a simulating environment of flue gas injection in the Xinjiang oil field. Materials
(Basel). 2018;11:1635.
22. Stone HL. Vertical, conformance in an alternating water-miscible gas flood. In: SPE Annual
technical conference and exhibition. Society of Petroleum Engineers; 1982.
23. Jenkins MK. An analytical model for water/gas miscible displacements. In: SPE Enhanced oil
recovery symposium. Society of Petroleum Engineers; 1984.
24. Pritchard DWL, Georgi DT, Hemingson P, Okazawa T. Reservoir surveillance impacts manage-
ment, of the Judy creek hydrocarbon miscible flood. In: SPE/DOE Enhanced oil recovery
symposium. Society of Petroleum Engineers; 1990.
25. Dawson AG, Jackson DD, Buskirk DL. Impact of solvent injection strategy and reservoir
description on hydrocarbon miscible EOR for the Prudhoe Bay Unit, Alaska. In: SPE Annual
technical conference and exhibition. Society of Petroleum Engineers; 1989.
26. Magruder JB, Stiles LH, Yelverton TD. Review of the means San Andres Unit CO2 tertiary
project. J Pet Technol. 1990;42:638–44.
27. Wylie P, Mohanty KK. Effect of wettability on oil recovery by near-miscible gas injection. In:
SPE/DOE Improved oil recovery symposium. Society of Petroleum Engineers; 1998.
28. Burger JE, Mohanty KK. Mass transfer from bypassed zones during gas injection. SPE Reserv
Eng. 1997;12:124–30.
29. Moissis DE, Wheeler MF, Miller CA. Simulation of miscible viscous fingering using a modi-
fied method of characteristics: effects of gravity and heterogeneity. SPE Adv Technol Ser.
1993;1:62–70.
30. Sahimi M, Rasaei MR, Haghighi M. Gas injection and fingering in porous media. In: Gas
transport in porous media. Springer; 2006, p. 133–68.
31. Nordbotten JM, Celia MA, Bachu S. Injection and storage of CO2 in deep saline aquifers:
analytical solution for CO2 plume evolution during injection. Transp Porous Media.
2005;58:339–60.
32. Juanes R, MacMinn C. Upscaling of capillary trapping under gravity override: application to
CO2 sequestration in aquifers. In: SPE improved oil recovery symposium. Society of Petroleum
Engineers; 2008.
33. Juanes R, MacMinn CW, Szulczewski ML. The footprint of the CO2 plume during carbon
dioxide storage in saline aquifers: storage efficiency for capillary trapping at the basin scale.
Transp Porous Media. 2010;82:19–30.
34. Abdelgawad KZ, Mahmoud MA. In-situ generation of CO2 to eliminate the problem of gravity
override in EOR of carbonate reservoirs. In: SPE Middle East Oil Gas show and conference.
Society of Petroleum Engineers; 2015.
35. Majidaie S, Khanifar A, Onur M, Tan IM. A simulation study of chemically enhanced water
alternating gas (CWAG) injection. In: SPE EOR conference at oil and gas West Asia. Society
of Petroleum Engineers; 2012.
36. Blaker T, Celius HK, Lie T, Martinsen HA, Rasmussen L, Vassenden F. Foam for gas mobility
control in the Snorre field: the FAWAG project. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1999.
37. Kovscek AR, Bertin HJ. Foam mobility in heterogeneous porous media. Transp Porous Media.
2003;52:17–35.
38. Haugen Å, Mani N, Svenningsen S, Brattekås B, Graue A, Ersland G, et al. Miscible and
immiscible foam injection for mobility control and EOR in fractured oil-wet carbonate rocks.
Transp Porous Media. 2014;104:109–31.
39. Stephenson DJ, Graham AG, Luhning RW. Mobility control experience in the Joffre Viking
miscible CO2 flood. SPE Reserv Eng. 1993;8:183–8.
40. Dellinger SE, Patton JT, Holbrook ST. CO2 mobility control. Soc Pet Eng J. 1984;24:191–6.
41. Sagir M, Tan IM, Mushtaq M, Pervaiz M, Tahir MS, Shahzad K. CO2 mobility control using
CO2 philic surfactant for enhanced oil recovery. J Pet Explor Prod Technol. 2016;6:401–7.
412 R. Azin et al.

42. Memon MK, Shuker MT, Elraies KA. Study of blended surfactants to generate stable foam in
presence of crude oil for gas mobility control. J Pet Explor Prod Technol. 2017;7:77–85.
43. Masalmeh SK, Wei L, Blom C. Mobility control for gas injection in heterogeneous carbonate
reservoirs: comparison of foams versus polymers. In: SPE middle east oil gas show and
conference. Society of Petroleum Engineers; 2011.
44. Rutqvist J, Tsang C-F. A study of caprock hydromechanical changes associated with CO2 -
injection into a brine formation. Environ Geol. 2002;42:296–305.
45. Moreno FJ, Chalaturnyk R, Jimenez J. Methodology for assessing integrity of bounding seals
(Wells and Caprock) for geological storage of CO2 . In: Greenhouse gas control technologies,
vol. 7. Elsevier; 2005. p. 731–9.
46. Jimenez JA, Chalaturnyk RJ. Are disused hydrocarbon reservoirs safe for geological storage
of CO2 ? In: International conference on greenhouse gas control technologies. Elsevier; 2003.
p. 471–6.
47. Lavrov A. Dynamics of stresses and fractures in reservoir and cap rock under production and
injection. Energy Procedia. 2016;86:381–90.
48. Saripalli KP, Mahasenan NM, Cook EM. Risk and hazard assessment for projects involving
the geological sequestration of CO2 . In: International conference on greenhouse gas control
technologies. Elsevier; 2003. p. 511–6.
49. Damen K, Faaij A, Turkenburg W. Health, safety and environmental risks of underground CO2
storage—overview of mechanisms and current knowledge. Clim Change. 2006;74:289–318.
50. Lucier A, Zoback M, Gupta N, Ramakrishnan TS. Geomechanical aspects of CO2 sequestration
in a deep saline reservoir in the Ohio River Valley region. Environ Geosci. 2006;13:85–103.
51. Rutqvist J, Birkholzer JT, Tsang C-F. Coupled reservoir–geomechanical analysis of the poten-
tial for tensile and shear failure associated with CO2 injection in multilayered reservoir–caprock
systems. Int J Rock Mech Min Sci. 2008;45:132–43.
52. Bohloli B, Skurtveit E, Grande L, Titlestad GO, Børresen M, Johnsen Ø, et al. Evaluation of
reservoir and cap-rock integrity for the Longyearbyen CO2 storage pilot based on laboratory
experiments and injection tests; 2014.
53. Preisig M, Prévost JH. Coupled multi-phase thermo-poromechanical effects. Case study: CO2
injection at In Salah, Algeria. Int J Greenh Gas Control 2011;5:1055–64.
54. Rutqvist J. The geomechanics of CO2 storage in deep sedimentary formations. Geotech Geol
Eng. 2012;30:525–51.
55. Kempka T, De Lucia M, Kühn M. Geomechanical integrity verification and mineral trapping
quantification for the Ketzin CO2 storage pilot site by coupled numerical simulations. Energy
Procedia. 2014;63:3330–8.
56. Okamoto I, Li X, Ohsumi T. Effect of supercritical CO2 as the organic solvent on cap rock
sealing performance for underground storage. Energy. 2005;30:2344–51.
57. Gaus I, Azaroual M, Czernichowski-Lauriol I. Reactive transport modelling of the impact of
CO2 injection on the clayey cap rock at Sleipner (North Sea). Chem Geol. 2005;217:319–37.
58. Carroll SA, McNab WW, Torres SC. Experimental study of cement-sandstone/shale-brine-CO2
interactions. Geochem Trans. 2011;12:9.
59. Smith SA, Sorensen JA, Steadman EN, Harju JA. Acid gas injection and monitoring at the Zama
oil field in Alberta, Canada: a case study in demonstration-scale carbon dioxide sequestration.
Energy Procedia. 2009;1:1981–8.
60. Smith SA, McLellan P, Hawkes C, Steadman EN, Harju JA. Geomechanical testing and
modeling of reservoir and cap rock integrity in an acid gas EOR/sequestration project, Zama,
Alberta, Canada. Energy Procedia. 2009;1:2169–76.
61. Bennion DB, Thomas FB, Bietz RF, Bennion DW. Water and hydrocarbon phase trapping in
porous media-diagnosis, prevention and treatment. J Can Pet Technol. 1996;35.
62. Brown JS, Al Kobaisi MS, Kazemi H. Compositional phase trapping in CO2 WAG simula-
tion. In: SPE Reservoir characterisation and simulation conference and exhibition. Society of
Petroleum Engineers; 2013.
63. Iglauer S, Paluszny A, Blunt MJ. Simultaneous oil recovery and residual gas storage: a pore-
level analysis using in situ X-ray micro-tomography. Fuel. 2013;103:905–14.
9 Challenges of Gas Injection 413

64. Zhou N, Matsumoto T, Hosokawa T, Suekane T. Pore-scale visualization of gas trapping in


porous media by X-ray CT scanning. Flow Meas Instrum. 2010;21:262–7.
65. Kimbrel EH, Herring AL, Armstrong RT, Lunati I, Bay BK, Wildenschild D. Experimental
characterization of nonwetting phase trapping and implications for geologic CO2 sequestration.
Int J Greenh Gas Control. 2015;42:1–15.
66. Thorne RJ. Transition to a low carbon economy; impacts to health and the environment. In:
Environmental determinants of human health. Springer; 2016. p. 169–201.
67. Mace MJ, Hendriks C, Coenraads R. Regulatory challenges to the implementation of carbon
capture and geological storage within the European Union under EU and international law. Int
J Greenh Gas Control. 2007;1:253–60.
68. van Egmond B. Developing a method to screen and rank geological CO2 storage sites on the
risk of leakage. NWS-E-2006-108, Utrecht University and TNO; 2006.
69. Oldenburg CM. Screening and ranking framework for geologic CO2 storage site selection on
the basis of health, safety, and environmental risk. Environ Geol. 2008;54:1687–94.
70. Li Q, Liu G, Liu X, Li X. Application of a health, safety, and environmental screening and
ranking framework to the Shenhua CCS project. Int J Greenh Gas Control. 2013;17:504–14.
71. Diao Y, Zhang S, Wang Y, Li X, Cao H. Short-term safety risk assessment of CO2 geological
storage projects in deep saline aquifers using the Shenhua CCS Demonstration Project as a
case study. Environ Earth Sci. 2015;73:7571–86.
72. Seligsohn D, Liu Y, Forbes S, Dongjie Z, West L. CCS in China: toward an environmental,
health, and safety regulatory framework; 2010.
73. Liu L-C, Li Q, Zhang J-T, Cao D. Toward a framework of environmental risk management
for CO2 geological storage in China: gaps and suggestions for future regulations. Mitig Adapt
Strateg Glob Chang. 2016;21:191–207.
74. Lee JS, Choi EC. CO2 leakage environmental damage cost—a CCS project in South Korea.
Renew Sustain Energy Rev. 2018;93:753–8.
75. Shuter D, Bilio M, Wilday J, Murray L, Whitbread R. Safety issues and research priorities for
CCS systems and infrastructure. Energy Procedia. 2011;4:2261–8.
76. Zhang L, Huang H, Wang Y, Ren B, Ren S, Chen G, et al. CO2 storage safety and leakage
monitoring in the CCS demonstration project of Jilin oilfield, China. Greenh Gases Sci Technol.
2014;4:425–39.
77. Smith SA, Sorensen JA, Steadman EN, Harju JA, Ryan D. Zama acid gas EOR, CO2
sequestration, and monitoring project. Energy Procedia. 2011;4:3957–64.
78. Evans DJ, Chadwick RA. Underground gas storage: an introduction and UK perspective. Geol
Soc Lond Spec Publ. 2009;313:1–11.
79. Li S, Dong M, Li Z, Huang S, Qing H, Nickel E. Gas breakthrough pressure for hydrocarbon
reservoir seal rocks: implications for the security of long-term CO2 storage in the Weyburn
field. Geofluids. 2005;5:326–34.
80. Li Z, Dong M, Li S, Huang S. CO2 sequestration in depleted oil and gas reservoirs—caprock
characterization and storage capacity. Energy Convers Manag. 2006;47:1372–82.
81. Reitenbach V, Ganzer L, Albrecht D, Hagemann B. Influence of added hydrogen on
underground gas storage: a review of key issues. Environ Earth Sci. 2015;73:6927–37.
82. Bai M, Shen A, Meng L, Zhu J, Song K. Well completion issues for underground gas storage
in oil and gas reservoirs in China. J Pet Sci Eng. 2018;171:584–91.
Chapter 10
Capillary Phase Trapping

Fatemeh Kazemi, Reza Azin, and Shahriar Osfouri

Abstract After primary and secondary recovery, substantial amounts of hydro-


carbon remain entrapped by capillary forces. This phenomenon, known as capil-
lary trapping. Different factors affect the microscopic trapping mechanism such as
pore structure, wettability, capillary pressure, interfacial tension, relative perme-
ability, initial saturation, and other properties of the rock and fluids. Understanding
the amount of trapped phase is vital for different applications such as enhance oil
recovery (EOR), enhanced gas recovery (EGR), and carbon capture and storage
(CCS). In EOR and EGR, the purpose is reducing residual oil saturation, while for
CCS, it is opposite, i.e. the target is maximising the amount of trapped CO2 . There
are two main capillary trapping mechanisms, snap-off and by-passing which are
described in detail in this chapter. The laboratory methods and empirical mathemat-
ical models that are used to measure and predict the trapped phase saturation are
discussed in this chapter. Furthermore, there are various techniques for the mobiliza-
tion of the trapped hydrocarbon phase. Gas injection is one of the effective methods
to remove the trapped phase. This chapter discusses some of the key aspects of phase
trapping and mitigation methods.

Nomenclature

OOIP Original Oil in Place

F. Kazemi · S. Osfouri
Department of Chemical Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: Fatemeh.kazemi@mehr.pgu.ac.ir
S. Osfouri
e-mail: Osfouri@pgu.ac.ir
R. Azin (B)
Department of Petroleum Engineering, Faculty of Petroleum, Gas and Petrochemical Engineering,
Persian Gulf University, Bushehr, Iran
e-mail: reza.azin@pgu.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 415
R. Azin and A. Izadpanahi (eds.), Fundamentals and Practical Aspects
of Gas Injection, Petroleum Engineering,
https://doi.org/10.1007/978-3-030-77200-0_10
416 F. Kazemi et al.

S Saturation
ϕ Porosity (%)
k Permeability (m2 )
ρ Density
CCS Carbon Capture and Storage
CO2 Carbon dioxide
N2 Nitrogen
scCO2 Supercritical carbon dioxide
gCO2 Gas carbon dioxide
EOR Enhanced Oil Recovery
Nc or Ca Capillary number
PV Pore Volume
IFT Interfacial Tension
CT Computed Tomography
USS Unsteady State Method
IR Initial-Residual
R2 Coefficient of determination
WAG Water alternating gas
N Number of data point
CDC Capillary desaturation curve
Ctrap Trapping capacity
S Saturation
Si Initial saturation
Sr Residual saturation
Snw Non-wetting phase saturation
Sw Wetting phase saturation
Snwi Initial non-wetting phase saturation
Snwr Residual non-wetting phase saturation
Sgi Initial gas saturation
Sgr Residual gas saturation
Soi Initial oil saturation
Sor Residual oil saturation
Swi Initial water saturation
Pc Capillary pressure
exp
Sor Experimental data of residual oil saturation
cal
Sor Calculated data of residual oil saturation
ex p
Sor Average of experimental data set of residual oil saturation
max
Snwr Maximum residual non-wetting saturation
max
Snwi Maximum initial non-wetting phase saturation
Snwc Critical non-wetting phase saturation
Kr Relative Permeability
Krw Wetting phase relative permeability
Krnw Non-wetting phase relative permeability
v Velocity (m/s)
10 Capillary Phase Trapping 417

μ Viscosity (Pa s)
σ Interfacial tension (mN/m)
γ Surface tension
θ Contact angle (degree)
N Bo Bond number
ρ Density difference
g Acceleration due to gravity
μi Invading fluid viscosity
μd Displacing fluid viscosity
rb Radius pore body
rt Radius pore throat
P Pressure difference
C Land trapping constant
R Radius
β Heterogeneity factor
L Length of the doublet
qw Water flow rate
v Velocity
μo Oil viscosity
μw Water viscosity
M Mobility ratio
α and β Spiteri et al. trapping constant
a and b Ma and Youngren trapping constant
Voi Volume of displaced water with oil during primary drainage
VDV Dead volume of cell
Vw f Volume of displaced oil during water flooding
Mdr y Mass of the dry core
Mdragnage Mass of the core after drainage
Mw f Mass of the core after water flooding

Subscripts

g Gas
o Oil
w Water
i Initial
r Residual or trapped
nw Non-wetting phase (oil or gas)
c Critical
418 F. Kazemi et al.

Superscripts

exp Experimental
cal Calculated
max Maximum

10.1 Introduction

At the end of primary and secondary hydrocarbon production by natural depletion and
conventional recovery, waterflooding or gas flooding, substantial amounts of valuable
hydrocarbon remain entrapped or bypassed due to the heterogeneity and capillary
forces of the reservoir. Up to 35–55% OOIP1 can be produced by conventional
methods (primary/secondary) and the remaining are the target of the EOR2 to increase
the recovery percentage [1]. For instance, the OOIP of a reservoir with an average
porosity 5%, 250 acer-ft bulk volume, and 20% connate water saturation, is 77,580
bbl. If the oil price is 50 $/bbl, the reservoir oil value will be $3,879,000. If we could
produce from this reservoir by 20% recovery factor, about 15,516 bbl of oil will be
produced with economic value of $775,800. About 62,064 bbl of OOIP get trapped
which corresponded to $3,103,200. This significant amount of oil is very important
since the consumption of energy is increasing every day and fossil fuels are expected
to remain the main energy resources in the future. So it is essential to maximize
the production of the hydrocarbon reservoir fluids using unconventional and new
technologies. Knowing the amount of the trapped phase and trapping mechanism is
crucial in proper designing of oil and gas recovery projects.
Furthermore, for mitigating greenhouse gas emission into the atmosphere, the
most promising technology is capillary trapping. It is the most favourable form
of a trapping mechanism of CO2 because of its storage security and capacity [2].
Carbon capture and storage process (CCS) is involved collecting CO2 from different
sources such as power plants, refineries, and industrial plants and storing it in deep
geological formations such as saline aquifers, depleted oil or gas fields, and coal
beds [3]. To maximize capillary trapping in CCS, a suitable approach should be
developed and implemented [4].
Therefore the volume of trapped phase saturation is a key factor in evaluating
recovery methods and also carbon capture and storage process [5, 6]. Furthermore,
aqueous phase trapping is a capillary imbibition effect which can take place in both
oil and gas reservoirs [7].

1 Original Oil in Place.


2 Enhance Oil Recovery.
10 Capillary Phase Trapping 419

10.2 Capillary Trapping

When a non-wetting phase displaced by a wetting phase, a portion of non-wetting


phase becomes immobilized because of capillary forces and is known as capillary
trapping [8, 9]. The amount of residual phase saturation does not reduce to zero even
if the capillary pressure approaches zero. Therefore, for example, in a water-wet oil
reservoir, the complete recovery of oil by imbibition process is impossible as shown in
Fig. 10.1. The same phenomena happen to the wetting phase, that complete removal
of the wetting phase from the porous medium by drainage process is impossible.
Various studies have been performed on capillary entrapment in water–oil, gas–
oil and three-phases, water–gas–oil systems. Different factors affect the microscopic
trapping mechanism are mentioned in Fig. 10.2 [9–11].
Among the various parameters that affect the trapping mechanisms, the structure
of cavities and wettability are two important factors in the immiscible displacement
of fluids in the reservoir and also capillary trapping. Although the structure of the
cavities cannot be changed much, the two other items i.e. fluid–fluid and fluid-rock
property can be changed by different methods such as surfactants flooding [12, 13],
gas injection [14], and wettability alteration [15–17]. The study of these parameters
and their contribution to capillary trapping is essential for understanding the displace-
ment process. Displacement of fluids in porous media has wide applications in natural
and industrial processes such as oil and gas recovery and geological carbon dioxide
storage. The basic and important phenomenon is the multiphase transport involves
the displacement of one fluid by another [18, 19]. The mechanisms of fluid displace-
ment in porous media are fundamentally different for drainage and imbibition [20].
The mechanisms of drainage and imbibition process, and therefore visualization of
fluids displacement and also phase trapping can be illustrated using different methods

Connate water
Grain

Low permeable pore channel


Oil
High permeable pore channel

Wetting phase (Water)

Trapped oil

Fig. 10.1 Illustration of oil trapping in water-wet oil reservoir after water flooding
420 F. Kazemi et al.

Density Interfacial
Viscosity ratio
difference tension

Fluid-fluid Relative
Wettability properties permeability
Capillary Initial
pressure saturation

Fluid-rock Pressure
interfacial gradient
properties and gravity

Properties
Pore of the
structure Trapping
fluids and
the flow

Fig. 10.2 Parameters affecting capillary trapping

such as transparent capillary glass plate networks. The simplest type of pore system
is the pore doublet which discusses in the future subsection.

10.2.1 Pore Structure

Porous media is referred to as consolidated such as limestone or sandstone, which


has become compacted over time or unconsolidated such as sand pack that are not
densely packed and have larger pore spaces for fluids to pass through [2].
It is known that in different geological formations, different quantities of hydro-
carbon can be trapped. It is showed that media with narrower pore throats, such as
sandstones, trap more non-wetting phase than media with wider pore throats [21].
Pore-scale petrophysical properties such as porosity, permeability, pore throat
radius, aspect ratio-the ratio of pore body diameter to pore throat diameter-, and
coordination number-the number of pore throats connected to each pore body-are
effective parameters in capillary trapping. For example, increasing aspect ratio leads
to increase capillary trapping [22] and the more coordination number, the greater
the chance that the non-wetting phase can be displaced during piston-like displace-
ment which is resulting in less residual trapping [23]. It is known that the porosity
has a significant impact on residual non-wetting saturations, but published studies
do not show consistent results as shown in Fig. 10.3. Most of the previous study
10 Capillary Phase Trapping 421

100
Kazemi et al., 2020 (n-Heptane/CO )
90 Kazemi et al., 2020 (n-Hexane/CO )
Tanino and blunt, 2012
80 Pentland et al., 2012
Wulling et al., 2009
70
Jerauld, 1996 (Fontainebleau)
60 Jerauld, 1996 (Frigg)
Declaud, 1991
Sor (%)

50 Aissaoui, 1983

40

30

20

10

0
0 10 20 30 40 50 60 70
φ (%)

Fig. 10.3 Experimental data on the effect of porosity on Snwr

reported an inverse relationship between porosity and residual saturation since in


high porosity the situation is not suitable for snap-off mechanisms [24–29]. On the
other hand, Pentland et al. experimentally demonstrated that the residual non-wetting
phase saturation increased against porosity increment in unconsolidated formations
and mentioned that the entrapment of a non-wetting phase is quantitatively different
between consolidated and unconsolidated media [21]. They also showed a weak
trend between permeability and residual saturation [21, 30].

10.2.2 Initial Saturation and Trapping Curve

Trapping curve or IR curve is the relationship between the initial trapped phase
saturation (Si ) and the residual trapped phase saturation (Sr ). Several studies have
discussed the trapping curve [9, 24, 29, 31–35]. These curves can be achieved by
Pc -Sw curves which contain drainage and imbibition curve. As shown in Fig. 10.4
applying different capillary pressures or different pore volume of non-wetting phase
injection, result in different initial oil saturations on the drainage curve. After imbi-
bition test, different corresponding remaining saturations can be achieved. Differ-
ence between the initial saturation and the remaining saturation represents capillary
pressure hysteresis.
The Trapping curve represented schematically in Fig. 10.5, correspond to the
points of Fig. 10.4.
422 F. Kazemi et al.

A
Capillary Pressure (Pc)

A1 Drainage
A2 A3

Imbibition B2 B3
B B1
-
0 Swc 0.2 0.4 0.6 0.8 1
Water Saturation (Sw)

Fig. 10.4 Pc -S curves which include the drainage capillary pressure ends at different values and
yields at different corresponding Sw during waterflooding

0.9

0.8

0.7

0.6
Snwr

0.5

0.4
Snwrmax

0.3
(A,B)
0.2 (A1,B1)
(A2,B2)
0.1
(A3,B3)
0
0 0.2 0.4 0.6 0.8 1
Snwimax
Snwi

Fig. 10.5 A schematic of capillary trapping curve for the water-wet porous media
10 Capillary Phase Trapping 423

Pentland et al., 2010 (Sandstone-Stainton) Pickell et al., 1966 (SandStone-Dalton1) Pickell et al., 1966 (Sandstone-Dalton2)
Pickell et al., 1966 (Sandstone-Dalton3) Pentland et al., 2010 (Sandstone-Berea) Pentland et al., 2010 (Sandstone-Clashach)
Masalmeh, 2002 (Sandstone) Pentland et al., 2011 (Sandstone) Al Mansoori et al., 2010 (Sandpack)
Pentland et al., 2010 (Sandpack) Tanino and Blunt, 2013 (Limestone-Indiana) Skauge et al., 2006 (Limestone)
Kazemi et al. 2020 (Sandpack) Kazemi et al. 2020 (Carbonate pack)

60

50

40
Sor (%)

30

20

10

0
0 10 20 30 40 50 60 70 80 90 100
Soi (%)

Fig. 10.6 Literature database of residual oil saturations versus initial oil saturations for water-wet
porous media

Figure 10.6 showed a selected literature survey of two-phase IR data, obtained


in water-wet media [9, 29, 30, 32, 34, 36–40]. All the studies show a monotonic
increase in the trapped phase saturation as a function of initial saturation. In water-
wet media, when the initial phase saturation is high, the amount of the trapped phase
is high, too, because there is more oil to trap and also in water-wet media, where the
contact angle is small, trapping is restricted by snap-off [5, 31]. In other words, higher
initial wetting phase saturations lead to a higher probability of the non-wetting phase
in larger channels, and therefore, the recovery is increased during imbibition. The
gradient of this trend decreases as the initial saturation increases in the studied water-
wet medium. It also seems that the trend of the IR curves was not linear. In other
words, the trend of the IR curves appeared linear up to a certain initial saturation,
and, after increasing linearly, at a specific initial saturation, the slope of the curves
decreased and then remained constant [29, 41].
There is scatter in the experimental data, due to different experimental conditions,
different contact angle and therefor the wettability, different rock types and pore
geometry, different interfacial tension (IFT), different fluid properties, and different
displacement mechanism [9, 24, 30, 31].
424 F. Kazemi et al.

10.2.3 Wettability

Wettability is the solid’s preference to be in contact with one fluid in a system of


two or more immiscible fluids [42]. In an immiscible two-phase flow system, one
fluid is considered wetting and the other non-wetting, i.e. one fluid having a stronger
attraction to adsorb to the surrounding solid phase than the other.
Reservoir rocks display a range of wettability, from being strongly water-wet that
spontaneously imbibe water, to strongly oil-wet or water repellent. Many rocks are
mixed-wet, with both water-wet and oil-wet pores and the other wettability state is
intermediate-wet where the rock appears to be largely non-wetting to both oil and
water [5, 43].
In oil reservoirs, some water is present that plays an important role in the displace-
ment process and oil recovery. In the absence of water, oil is always wetting phase
and gas is non-wetting phase. While in the presence of water, depending on rock
wettability, oil could be wetting, non-wetting or partially wetting phase. Each one
of these conditions could have significant effect on displacement process [44]. The
distribution of the wetting and non-wetting fluids is determined by the pore geom-
etry and the medium’s wettability. The wettability of a solid show how fluids move
throughout the pore space. For instance, in water-wet systems, the wetting fluid will
coat the walls of the solid and prefer to occupy and invade small crevices and pore
throats, while non-wetting fluid will tend to invade and occupy the larger pores and
reside in the center of pore bodies [20].
Clean reservoir rocks or minerals (quartz or calcite) that considered for geological
sequestration of atmospheric gases (often saline aquifers) are considered water-wet.
Once due to oil migration, the rock’s surface stays in contact with crude oil for
sufficient time, by making an oil layer on the surface of the rock, the wettability is
changed to more oil-wet since the heavy polar components of the crude oil (resins
and asphaltene) precipitate on the surface of the rock. Besides, the composition and
pH of the brine, and composition of the crude oil have significant impact on the
wettability alteration process [45, 46]. Hence, most of the oil reservoirs are believed
to be mixed-wet or intermediate-wet.
Figure 10.7 shows the possible configuration of oil and water saturations in water-
wet, mixed-wet and oil-wet porous media. Contact angles are often reported rela-
tive to the aqueous phase. If the contact angle ranged 0–70°, the solid surface is
hydrophilic and is termed water-wet medium. A surface with no preference for
either fluid that the contact angle ranged 75–105°, is termed intermediate-wet. If
the contact angle ranged 105–180°, the solid surface is hydrophobic and depending
on the second fluid present in the system, the medium is termed oil-wet or gas-wet
[47].
The residual oil saturation in oil-wet and mixed-wet media is lower than water-wet
ones. The reason is the oil layers maintain connectivity. For intermediate-wet rocks,
residual saturation can also be low since snap-off is suppressed [5].
Understanding wettability and its influence on capillary trapping is a key factor
in estimating storage and recovery efficiencies.
10 Capillary Phase Trapping 425

Oil
Water

(a) Water-wet (b) Mixed-wet (c) Oil-wet

Fig. 10.7 Water and oil distribution in Different wettability. a Oil remains in the center of the large
pores. b Oil has displaced water from some surfaces but still trapped in the center of the water-wet
regions. c Water remains in the center of the larger pores and the oil surrounds the water

10.2.4 Relative Permeability

Relative permeability is the ratio of the effective permeability, the permeability of


one fluid in the presence of another fluid, to the absolute permeability. Three main
factors include pore geometry, fluid saturation, and wetting state affecting relative
permeability. For example in a water-wet system, the water relative permeability
is lower than the oil relative permeability since the water fills the narrow pores in
contrast, the oil fills the large pores, giving the poor connectivity of water [48].
In water-wet system, the residual oil saturation is high, since oil can be trapped
in the centre of the larger pore by snap-off, which is discussed further in the next
subsection. For an intermediate-wet system, both oil and water have comparable
relative permeabilities since the surface has an equal tendency to be coated by either
fluid. In an oil-wet system, oil has a lower relative permeability than water since oil
fills the narrow pores and water fill the large pores giving low flow conductance to
oil. However, the oil remains connected in layers until a low residual saturation is
reached. For a mixed-wet system, the behaviour is characterised by low oil and water
relative permeabilities and so low residual oil saturation. This can be explained by the
low conductance of oil layers, and because it is hard for water to flow across different
wettability fractions [49]. Therefore relative permeability is essential to understand
oil recovery and thus plan proper injection schemes.
Laboratory investigation showed that non-wetting phase relative permeability
follows distinct path since the direction of saturation change is altered from drainage
to imbibition, and this path is a function of residual non-wetting phase satura-
tion, which depends on the initial saturation established in the drainage direction
(Fig. 10.8) [50]. So relative non-wetting phase permeability is also a function of the
saturation history of the medium [11, 51].
426 F. Kazemi et al.

Fig. 10.8 Typical relative


100 Snw 0
permeability-saturation
curves for a porous medium 1.0

krnw

kr krw

0
0 Swr Sw Snwr 100

10.2.5 Interfacial Tension

A fluid phase in contact with another fluid has free interfacial energy between the
phases caused by the net attractive forces that exist in those molecules that are near
the interface. Molecules at the interface of two immiscible fluids are attracted to
molecules of its own phase by a force more strongly than the force attracting it to
molecules of the opposite phase across the interface. The cohesive forces acting
on a molecule inside a fluid phase, and a molecule at the interface (liquid–gas or
liquid–liquid) between fluid phases are shown in Fig. 10.9 [51].
When this force encountered between two immiscible fluids, it is called the inter-
facial tension (IFT), σ [J/m2 or N/m]. Between liquid and a gaseous phase and solid

Fig. 10.9 Forces on


molecules of a liquid Immiscible Fluid 2

Fluid-Fluid interface

Fluid 1
10 Capillary Phase Trapping 427

surface, this same force is called the surface tension, γ, which related to the wetta-
bility. Interfacial tension is zero between completely miscible fluids, and it is constant
for any pair of phases [51, 52].
The decrease of interfacial tension between fluid phases leads to a decrease in the
magnitude of the capillary pressure for a fixed radius of curvature and wettability
condition. The interfacial tension needs to be large for the capillary forces to be
dominant. If the IFT is weak, viscous forces will be dominating. Viscous forces are
the force created by a difference in the fluids’ viscosity; it is also normally referred
to as Darcy force [53].
The capillary forces vary linearly with the interfacial tension and depend on the
radius of the capillary tube. Therefore, when the interfacial tension is high, the
capillary forces are high, too. In high capillary forces (insignificant viscous force
or low flow rate) the interface in the small tube reaches the downstream faster, and
trapping happens in the large tube. Then, the trapped phase saturation increase.
However, it should be noted that since the viscosity of the fluid affects the residual
saturation, finding the exact effect of the IFT is impossible [29].

10.2.6 Capillary, Viscous and Gravitational Forces

On a microscopic scale, capillary forces play a very important role in the displacement
of one immiscible fluid by another immiscible fluid. Also there are important in
determining the amount of trapped or residual phase saturation in either laboratory
or field displacements. Capillary forces, combined with the spatial variation of rock
properties in an oil reservoir, may reduce the recovery factor of the reservoir. For
instance, large oil quantities are difficult to remove from parts of reservoirs with
small scale heterogeneities [54]. The pore structure of the porous medium governs
the fluid–fluid interfaces established in the porous medium. Entrapment of one fluid
by another immiscible fluid during drainage and imbibition displacement is the net
result of the interaction of capillary forces at the microscopic pore level. The effects of
capillarity on immiscible two-phase flow in heterogeneous porous media have been
studied by several authors [55, 56]. In systems where capillary forces and interfacial
interactions dominate flow processes, in drainage, the non-wetting phase invades
the largest available pore throats. It progressively invades smaller and smaller pore
throats, and non-wetting phase pressure increases relative to wetting phase pressure
and drainage proceeds [10].
The phase diagrams which compare the relative importance of gravity, viscosity,
and capillarity have been developed by many previous researchers which illustrated
in Fig. 10.10. This diagram indicates which flow regime will dominate, following
the original phase diagram proposed by Lenormand et al. [10, 57, 58].
428 F. Kazemi et al.

0
Stable
Displacement

-2

Viscous
Fingering
-4
log (Ca)

-6

Capillary
-8 Fingering

-10
-6 -4 -2 0 2 4
log (M)

Fig. 10.10 Classification of drainage flow patterns based on the Capillary Number (Ca) and
viscosity ratio (M) of the fluid displacement experiments. Solid lines indicate the flow regime
boundaries as originally defined by Lenormand et al. [10]; dashed lines indicate boundaries reported
by Zhang et al. [58]

10.3 Essential Dimensionless Groups in Phase Trapping

In porous media, immiscible fluids displacement affected by capillary, viscous


and buoyancy forces. Dimensionless groups describe the relative influence of the
governing forces on the flow regime that allow us to compare the results in similar
geometry and help to reduce the number of parameters that are needed to be consid-
ered in the problem statement and therefore decreasing the experimental effort
required. It has been demonstrated that the volume of the trapped phase can be
correlated to the magnitude of dimensionless groups [59–61].

10.3.1 Capillary Number

Two-phase flow dynamics is typically characterized by a dimensionless group termed


capillary number (Nca ) that is used to describe the ratio of viscous forces to capillary
10 Capillary Phase Trapping 429

forces. The most common formula for the capillary number is defined as Eq. 10.1.

V iscous For ces νμ


NCa = = (10.1)
Capillar y For ces σ cos θ

where v is the Darcy injection velocity of the invading fluid in cm/s, μ is the viscosity
of the invading fluid in g/cm.s, σ is the interfacial tension between the invading and
displaced fluids in dyne/cm, and θ is the contact angle in degree [60, 62–64]. The role
of the invading or displaced fluid changes between imbibition and drainage events
[2].
The residual non-wetting phase saturation is dependent on the capillary number
[10, 23, 58, 63, 65]. As capillary number increases (viscous forces dominate the
capillary force), residual phase trapping decreases. A dramatic decrease in residual
phase saturation (depending of the kind of non-wetting phase) is observed around a
critical capillary number of 10–5 to 10–8 . It is attributed to the suppression of snap-off
at larger N ca [66]. Viscous forces suppress snap-off and mobilize ganglia, leading
to less trapping [5]. So in experimental study, the capillary number is measured to
ensure that the condition is capillary dominated and the capillary trapping is expected
to happen.
Dependence of residual saturation on the capillary number is often illustrated by
the capillary desaturation curve (CDC). This curve, shown in Fig. 10.11, indicates that
Residual non-wetting or wetting phase saturation (%)

Snwr
30
Swr

20

10 Non-wetting
critical Ca Wetting
critical Ca

0
-7 -6 -5 -4 -3 -2
10 10 10 10 10 10
Nc

Fig. 10.11 Capillary desaturation curve, inversely trend of residual saturation with the capillary
number [68]
430 F. Kazemi et al.

70
n-Hexane/CO
60 n-Heptane/CO
n-Hexane/CH
50 n-Heptane/CH
n-Hexane/N
40 n-Heptane/N
Sor

30

20

10

0
0 0.00001 0.00002 0.00003 0.00004 0.00005 0.00006
Nc

Fig. 10.12 Different Capillary desaturation curves of the gas–oil system in unconsolidated sand
pack [29]

residual saturation decreases with the increasing capillary number [67] and shows
the critical capillary number (Nc ) that is the number, at which the residual saturation
begins to decline.
Ding and Kantzas conducted two-phase experiments on Berea sandstone and
determined critical capillary numbers of the non-wetting phase is 10–5 for water–oil
and 10–8 for water–gas systems [69]. Shen et al. reported the non-wetting critical
capillary numbers of sandstone is 10–3 that was several orders of magnitude higher
than that reported by Ding and Kantzas. They suggested that porous media charac-
teristics have a strong influence on capillary trapping behaviour [70]. The capillary
number in consolidated media is one or two orders of magnitude lower than uncon-
solidated media. However, the trend of capillary number is the same [71]. Kazemi
et al. studied the effect of capillary number on residual oil saturation in the gas–oil
system. As shown in Fig. 10.12, the trapped oil phase increased as the capillary
number increased since the viscous force increase by increasing capillary number,
and the condition is not favourable for capillary trapping.

10.3.2 Bond Number

In a three-dimensional system, gravity forces must be taken into consideration as well


as viscous and capillary forces. The bond number affects more in vertical displace-
ment. The effects of gravity on the system can be quantified by the bond number,
which indicates the relative importance of gravitational forces to capillary forces, as
reported in Eq. 10.2.

Gravit y For ces ρgk


N Bo = = (10.2)
Capillar y For ces σ
10 Capillary Phase Trapping 431

where ρ is the density difference between the invading and displaced phases
[kg/m3 ], g is the acceleration due to gravity [9.8 m/s2 ], k is the absolute perme-
ability [m2 ], and σ is the interfacial tension between the invading and displaced
fluids [N/m]. Typically in reservoir displacements N Bo << 1. It is demonstrated that
if N Bo increased, i.e. gravitational forces dominated over capillary forces, the trapped
phase saturation decreased [66, 72].

10.3.3 Viscosity Ratio (Mobility)

The mobility ratio, M, is often used as a method of comparing the relative effect of
the viscosities on trapped phases in porous media. There are various definitions of
mobility ratio. The most common Equation is:
μi
M= (10.3)
μd

where μi represents the viscosity of the invading fluid and μd represents the viscosity
of the displaced fluid [10, 58].
Capillary numbers are often plotted against mobility ratios as a method of visu-
alizing the dependence of viscous forces to capillary forces. Based on the resulting
plot, three distinct regimes emerge and are defined as viscous fingering, capillary
fingering, and stable displacement regimes [10]. More detail discussed in the prior
subsection.

10.4 Importance of Trapping

10.4.1 Oil Recovery

Capillary trapping due to the reservoir heterogeneity plays an important role in many
oil recovery processes where the multiphase flow is encountered. In oil reservoirs,
we design the process to minimize the capillary trapping of oil, hence maximizing
recovery. Oil production has three stages: primary, secondary, and tertiary. Primary
recovery is the process when the oil comes out of the reservoir to the production well
due to the natural depletion of the reservoir pressure. Secondary recovery comes after
primary recovery, and here water and/or gas are injected to flood the oil out of the
reservoir. Tertiary recovery comes after secondary recovery, and the aim is to recover
more oil after secondary recovery. The fluids injected in tertiary recovery are usually,
surfactants, polymers, or CO2 (called EOR) [73]. CO2 can be injected in depleted
oil reservoirs to displace the oil, while CO2 remains in the reservoir. Besides, this
process will maintain reservoir pressure and production rates [74]. In CO2 enhance
432 F. Kazemi et al.

oil recovery, storage occurs as CO2 displaces hydrocarbons and the injected CO2
is trapped within the reservoir’s pore spaces through capillary forces and the other
mechanisms that will be discussed later. CO2 EOR projects can be converted to CCS
projects at the end of their operating lifetimes [75].
The trapped phase saturation and mobilization of residual saturations are very
important in many aspects of enhanced recovery [76–79]. Trapping mechanisms
determine what proportion of oil is recovered and, consequently, how much oil
remains for possible recovery by tertiary methods. The volume of oil remaining
in swept zones and the fraction of this oil that is recovered, are critical economic
factors in the application of enhanced recovery processes [11].
Knowing the amount of the trapped phase and mechanisms of trapping is crucial
in proper designing of oil recovery projects.

10.4.2 CO2 Storage

Climate change is one of the major challenges facing humankind. We need to capture
and store CO2 that is emitted by fossil-fuel and industrial plants. The CO2 is trapped
by dissolution, mineralization, structural and stratigraphic confinement, and capillary
trapping. Among all CO2 trapping mechanism, Capillary trapping is a rapid, reliable
and effective mechanism to ensure safe storage in the subsurface [24, 80–86]. CO2
trapping occurs when brine displaces and traps the CO2 as discontinuous droplets or
through brine injection [18, 30, 81, 83, 86, 87]. Primary drainage capillary pressure
that shows the threshold pressure to enter the pore, and the amount of trapped CO2 as
a function of initial CO2 saturation which named the trapping curve, that determine
how much CO2 can be rendered immobile by capillary forces, are two key factors
that affect the subsurface behaviour of CO2 [83].
In CCS (Carbon Capture and Storage) applications the fraction of the rock volume 
that contains trapped phase termed capillary trapping capacity Ctrap = φ Snwr ,
which is the product of residual saturation and porosity, is of interest since it states
how much CO2 can be stored securely per unit rock volume [24, 30].
There are several geological sites such as saline aquifers and oil reservoirs that
CO2 can be stored. Saline aquifers provide the largest potential for storage. To ensure
security and prevent leakage, the physical and chemical mechanisms of CO2 capillary
trapping must be understood [2, 30, 88, 89]. As illustrated in Fig. 10.13, many
researchers studied capillary trapping of CO2 and other gases like nitrogen and air
[27, 33, 35, 39, 41, 90–93].
Studies on Carbon Capture and Storage (CCS) and Improved Oil Recovery (IOR)
have different emphasis; in CCS the success of the project depends on its ability
to trap CO2 in the subsurface and increases storage security to prevent it from
migrating upwards, while for IOR trapping phenomena limits oil recovery therefore
the objective in oil recovery methods is to minimize trapped oil [31].
In CCS, three displacement process take place: primary imbibition, primary
drainage and secondary imbibition. The purpose of the first stage in an experimental
10 Capillary Phase Trapping 433

Ma and Youngren, 1994 (Sandstone) Crowell et al., 1966 (Sandstone)


Pentland et al., 2011 (Sandstone) Kralik et al., 2000 (Sandstone-A3)
Kralik et al., 2000 (Sandstone-A4) Kralik et al., 2000 (Sandstone-C1)
Kralik et al., 2000 (Sandstone-C3) Kralik et al., 2000 (Sandstone-C4)
El-Maghraby and Blunt, 2012 (Sandstone-Berea) Krevor et al., 2011 (Sandstone)
Akbarabadi and Piri, 2013 (Sandstone-Berea) Akbarabadi and Piri, 2013 (Sandstone-Berea)
Akbarabadi and Piri, 2013 (Sandstone-Nugget) Jerauld, 1997b (Sand Stone-Prudhoe Bay)
Al Mansoori et al., 2010 (Sandpack) El-Maghraby and Blunt, 2012 (Limestone-Indiana)
50

45

40

35

30
Sgr(%)

25

20

15

10

0
0 10 20 30 40 50 60 70 80 90
Sgi(%)

Fig. 10.13 Literature database of residual gas saturation versus initial gas saturations for water-wet
porous media

study is to achieve a condition like a saturated aquifer that is the site used to CO2
sequestration. Then, during the primary drainage, the CO2 is injected into the porous
media. In this process, part of the wetting phase (water) will remain trapped in the
pores as irreducible wetting fluid (Sirr ). According to Laplace’s Law, wetting phase
trapping will happen in the smallest pores because the capillary forces required to
remove the fluid from small pores is too large in comparison with larger ones. There-
fore wetting phase is displaced from the larger pores first since they have lower capil-
lary pressure threshold. The drainage process will continue until drainage is forcibly
stopped or until the system reaches to an equilibrium condition. At the final stage,
the water reimbibed into the pore space and preferentially fills the smaller pores first
because of the reduced capillary pressure requirements. During this process, the non-
wetting phase may become surrounded by the wetting phase, since there is pressure
difference between the wetting and non-wetting phases. So doublet is disconnected
from the rest of its fluid phase, and can’t leave the pore space [2].
One of the complications of CCS is that both the drainage (CO2 injection) and the
imbibition (water flooding) processes should be manipulated; while in oil recovery
the non-aqueous phase is initially present in the medium and only the imbibition
process should be engineered [66].
434 F. Kazemi et al.

10.4.3 Condensate Recovery

Gas condensate is a kind of natural gas reservoir that the liquid and gas phases coexist
as the reservoir pressure drops below its dew point. The retrograde behaviour of gas
condensate reservoir lead to liquid accumulation, especially near the wellbore, and
therefore, this phenomena significantly restricts gas inflow and reduces the produc-
tion of the reservoir [94]. When condensate liquid forms in a gas reservoir, it is
immobile because of capillary forces acting on the fluids. On the other word as the
condensate is formed, the capillary forces start to overcome the viscous forces of the
gas, causing trapping in the pores. The reduction of mobility near-wellbore area is
consequence of liquid trapped in the pores of the rock when the pressure difference
between the non-wetting and the wetting phase is large [95].
Microscopic liquid droplet tends to be trapped in small pores or pore throats.
The condensate that forms in most of the reservoir is lost to production unless the
depletion plan includes gas cycling. Even in a lean gas reservoir with low liquid
drop-out, condensate blockage can be significant, since capillary forces can retain a
condensate that builds to a high saturation over time [96]. Capillary force opposes
dropped out liquid phase from flowing into the wellbore in a liquid (condensate) wet
rock. Therefore it leads to high condensate saturation and decreases the gas phase
relative permeability and so the productivity of the well. At the bottom hole and
in high permeability reservoirs, we deal with high velocity at the wellbore, which
lessens the trapped oil saturation [95]. As capillary number increases, the wetting
and non-wetting relative permeabilities increase, given that the fluid viscous forces
overcome the capillary forces, decreasing trapping of liquids [97].
Several observations and studies indicate condensate liquid drop-out and hydro-
carbon trapping in reservoirs, especially in the near-wellbore regions [98, 99]. Main-
taining the pressure of a wellbore above the dew point can prevent the formation
of a liquid. The classical remediation techniques for dealing with this phenomenon
include revaporization through gas injection [100–102]. Gas cycling is a method that
have been proposed to restore gas production rates after condensate blockage. Injec-
tion of dry gas lead to vaporization of condensate and increases its dew point pres-
sure [103]. Carbon dioxide and nitrogen injection have been investigated to improve
gas condensate recovery [104–108]. Jamaluddin et al. experimentally showed that
propane decreases the dew point and vaporize condensate more efficiently than
carbon dioxide [109]. Chemical and mechanical method is used to enhance the
productivity of the well. For instance, hydraulic fracturing has been used to enhance
gas productivity, but it is not always feasible or cost-effective. Hydraulic fracturing
can increase the bottom-hole pressure and so successfully restored the gas produc-
tivity of a well [110, 111]. Furthermore, solvents were used to increase gas rela-
tive permeability near the well-bore due to condensate and water blocking [112].
Recently, novel techniques for the wettability alteration of near-wellbore formation
have been suggested to reduce liquid trapping in gas condensate reservoirs [99, 101,
113–119]. Wettability alteration is the only permanent method used for this purpose.
10 Capillary Phase Trapping 435

10.5 Capillary Trapping Mechanisms

There are two major mechanisms for capillary trapping: snap-off and by-passing
[120–122]. These mechanisms describe trapping based on pore geometry and
wettability.

10.5.1 Snap-off

Snap-off is an immiscible phase entrapment mechanism that occurs when the capil-
lary pressure within throats exceeds associated pores [123]. Oh and Slattery [124]
and Chatzis et al. [125] were the first ones who used sinusoidal geometry for theoret-
ical and experimental investigation. It was determined that almost 80% of the trapped
phase accrue in snap-off geometry and 20% in pore doublet and the combination of
both geometries [125].
Snap-off occurs as non-wetting fluid in a pore is displaced from a pore body into a
pore throat. Capillary trapping is a function of contact angle as well as pore geometry.
The combined effect of contact angle and pore geometry control the curvature of a
fluid–fluid interface and determine the potential for snap-off. Therefore the mech-
anism strongly depends on wettability and the aspect ratio-the ratio of pore body
diameter to pore throat diameter (Eq. 10.4)—in a porous media.

por e body diameter rb


Aspect ratio = = (10.4)
por e thr oat diameter rt

This phenomenon is usually called snap-off, choke-off, pinch-off and break-up.


The mechanism happens in this way: consider a medium with smooth and water-wet
walls that has a high aspect ratio which the pore bodies are much bigger than the
pore throats. When water displacing a non-aqueous liquid in this condition a thin
layer of water phase reaches the downstream of the pore throat, and hence a large
blob of organic liquid remains in the pore body and disconnected from bulk of the
liquid, and a single blob is trapped by snap-off in each pore as shown in Fig. 10.14a.
In low aspect ratio, i.e. the pore throats are almost in the same size of the pore bodies,
as depicted in Fig. 10.14b, the fluid can be completely displaced, and no trapping
happen. It is important to note that the large aspect ratio leads to extensive trapping
under water-wet condition, while it has little effect on trapping under intermediate
wettability [126].
The effect of wettability showed in Fig. 10.15. When the invading fluid that is
strongly wetting (the contact angle is near 10°), reaches the downstream of the throat,
a large blob of displaced phase (oil or gas) remaining in the pore body and trapped
which is indicated in Fig. 10.15a. When the interface with a contact angle of about
90° passing from a pore throat through a high aspect ratio pore body with smooth
walls, causes the curvature of the interface to remain relatively small and therefore
436 F. Kazemi et al.

Invading fluid Invaded fluid

Invading fluid Invaded fluid

Fig. 10.14 Effect of aspect ratio on non-aqueous phase trapping, a High aspect ratio and snap-off
capillary trapping. b Low aspect ratio and no snap-off trapping [126]

Contact angle ~ 5 Contact angle ~ 90º

Snap-off
Pore throat Pore throat
Invading fluid
Advancing Displaced Advancing Displaced
phase phase phase phase
Pore body Pore body
a b

Fig. 10.15 Displacement of two immiscible fluids in a circular, smooth and high aspect ratio pore.
a Intermediate wetting conditions. b Strongly wetting conditions [126]

when the interface reaches the downstream, a small portion of invaded phase (oil or
gas) remains in the pore, and no trapping occurs (Fig. 10.15b).
So snap off only occurs in certain cases:
1. Strongly water-wet, i.e. the rock needs a wetting film completely separating
the rock from the non-wetting phase. Therefore snap-off is facilitated when the
contact angle is small.
2. High aspect ratio, i.e. large pore bodies with very small pore throats. At a high
porosity, the situation is not suitable for snap-off mechanism [24]. In other
words, in a unconsolidated or high porosity media, there are usually well-
connected pores and less contrast in the size of wider pores and narrower throats;
hence, less suitable conditions for snap-off.
It should be determined whether the flow is dominated by snap-off or by frontal
displacement. Increasing the flow rate led to less trapping since snap-off is favoured
by low flow rates [127].
10 Capillary Phase Trapping 437

10.5.2 By-Passing

Pore doublet, as a model of reservoir rock, is used to describe the by-passing trapping
mechanism. The pore doublet model is supposed to show the balance between viscous
and capillary forces. As illustrated in Fig. 10.16, a pore doublet consists of two parallel
tubes with different diameters connected on both ends [128].
The wetting phase and non-wetting phase tend to invade the smaller and the larger
pores, respectively. This leads to a special trapping mechanism known as by-passing.
The assumptions of this model are as follow:
1. Well-developed Poiseuille flow occurs in each path of the doublet.
2. Presence of the interface does not affect the flow.
3. The wetting and the non-wetting phase have equal viscosity.
4. When the wetting or non-wetting interface reaches the downstream in either
capillary, resident fluid trapped in another capillary.
Assumption 1 and 2 would be accurate if the doublet length were larger than the
largest capillary radius and the flow rate were very slow. Based on these assumptions,
we can use Eq. 10.5 [5]. To investigate the trapping behaviour of the pore doublet
model, the ratio of the average velocity in each of the capillary achieved using
Eq. 10.5.
 
v2 4Ncap + −1
1
β
=   (10.5)
v1 4Ncap
− β 2 β1 − 1
β2

where β = RR21 is the heterogeneity factor and Ncap = π RμLq w


is a dimensionless
1 σ cos θ
3

ratio of viscous to capillary forces which is named local capillary number.


In large flow rate (viscous forces dominance capillary forces), the N cap is high
and the second terms of the numerator and denominator expressions are negligible.
So the velocity in each path of the doublet is proportional to its squared radius.
 2
v2 R2
= β2 = (10.6)
v1 R1

Wetting phase R1

R2 Non-wetting phase

ΔP

Fig. 10.16 Schematic of the pore doublet structure


438 F. Kazemi et al.

Therefore the interface in the large capillary of the pore doublet will reach the
downstream before the interface in the small path, and the non-wetting phase will
be trapped in smaller tube. To explain how a non-wetting phase can be trapped in
a large flow rate, a typical value of parameter involve Eq. 10.3, are assumed. R1 =
5 μm, R2 = 25 μm, L = 50 μm, μo = μw = 1 cp, σ = 20 dyne/cm, θ = 10° and qw =
5 cc/min. In this case, the viscosity ratio will be 20.28, which is shown the interface
in larger capillary move faster and reach the down steam sooner that the interface in
smaller one. Therefore the remaining fluid in the small path will be trapped.
If the viscous forces are negligible compare to capillary forces, the small capillary
imbibe fluid faster than the supplied rate at the inlet of the doublet. The interface
velocity in the large radius path is negative in the fluid-starved doublet and at the
same time, the velocity in the small path is greater than that at the doublet inlet. For
example, if the water flow rate is 0.001 cc/min with above value of the parameters,
the displacement in two capillaries is counter-current. The flow rate in small and
large capillary are 1.85E−3 and −0.85E−3 cc/min respectively, and the velocity
ratio will be −0.018. It shows that the interface seals off the small capillary at the
inlet and the flow in the small one will be zero. This condition is in disagreement with
the derivation assumption that was the interface in two branches of the pore doublet
advanced simultaneously in a co-current pattern. Since visualization of negligible
viscous force is hard, we can imagine an intermediate case that viscous forces are
small but not negligible. In this case, the doublet is no longer starved for fluid, but
the interface in the small capillary is still faster than in large one. In the mentioned
example, if the water flow rate is in the rage of 2E-3 and 5E-3 cc/min, this condition
would take place, and trapping of the non-wetting phase will happen in the larger
capillary. Most flow in permeable media will be approximate by this intermediate
case. So in low flow rates, we expect more trapped phase saturation compare to high
flow rates.
Figure 10.17 showed velocity ratio versus water flow rate to know the effect of
flow rate on the interface movement and trapping in a considered pore doublet.
To investigate the effect of wettability on pore doublet mechanism, the velocity
ratio versus contact angle was plotted. Velocity ratio formula is different in the imbi-
bition and the drainage processes. In oil-wet media, the waterflooding is a drainage
process and the pressure drop across each capillary is the sum of the capillary pressure
and the viscous pressure drop as Eq. 10.6.

P1 + Pc1 = P2 + Pc2 (10.7)

So the velocity ratio expression for drainage process is as follow:


 
v2 4Ncap − 1
β
−1
=   (10.8)
v1 4Ncap
+ β2 ∗ 1
−1
β2 β
10 Capillary Phase Trapping 439

2.5

1.5
v2/v1

0.5

0
0.00E+00 2.00E-02 4.00E-02 6.00E-02 8.00E-02 1.00E-01 1.20E-01
-0.5 qw (cc/min)

Fig. 10.17 Effect of water flow rate on the velocity of the interface in two branches of the pore
doublet

water-wet, qw=0.0005
9 water-wet, qw=0.005
8 water-wet, qw=0.01
oil-wet, qw=0.0005
7 oil-wet, qw=0.005
6 oil-wet, qw=0.01
5
v2/v1

4
3
2
1
0
-15 5 25 45 65 85 105 125 145 165 185
-1
θ(°)

Fig. 10.18 Effect of water contact angle on the velocity of the interface in two branches of the pore
doublet

As shown in Fig. 10.18, by increasing the contact angle from 0° to 90° the velocity
ratio increased and therefore the interface in the large path, reaches the downstream
sooner and trapping take place in smaller tube. In oil-wet media, by changing the
wettability toward strongly oil-wet, the velocity ratio decreased until the velocity
of the interface increased in smaller capillary and trapping take place in larger one.
There is probably an optimum velocity ratio that the least trapping will happen.
It is important to note that trapping estimation using pore doublet greatly
overestimate residual saturation at low capillary number.
In the drainage process, the main trapping mechanism is by-passing although
during the imbibition process the non-wetting phase could be trapped by by-passing
mechanism too. The smaller throat is swept first during the imbibition because of
440 F. Kazemi et al.

Non-wetting phase

Wetting phase

Fig. 10.19 Trapping during the a Drainage and b Imbibition process in the pore doublet model

higher capillary pressure. So the trapping of the non-wetting phase will happen in
the larger pore (Fig. 10.19a). Furthermore, larger throat is invaded first during the
drainage process and the wetting phase trapped in smaller one (Fig. 10.19b). It should
be noted that the pore walls are smooth and strongly water-wet [76].
This kind of trapping tends to occur in poorly sorted media or rock with dual-
porosity networks [129].

10.6 Mobilization of Trapped Phase

About 45% of the OOIP can be produced by conventional primary and secondary
recovery techniques [130]. The rest of oil that is estimated by The U.S Department
of Energy about 377 billion barrels of discovered oil, is still trapped in reservoir
pores [131]. Due to the increasing energy consumption in the world and deple-
tion of the conventional hydrocarbon reservoir, there is need extremely to produce
more hydrocarbon fluids and therefore, new technologies for increasing recovery and
productivity of the reserve.
There are various techniques for the mobilization and removal of the hydrocarbon
phase and also aqueous phase traps. Any process which increases oil production after
primary and secondary recovery by decreasing the residual oil saturation is defined
as the EOR method. In this process, different fluids that are not normally present in
the reservoir like surfactants, polymers, foams, and various gases, are injected to the
reservoir. Then the fluids interact with oil/brine/rock of the porous media to increase
oil production [132]. Indeed in the tertiary recovery of the remaining phase, an addi-
tional energy source is required to microscopically and macroscopically mobilization
and displacing residual and by-passed phase. Various methods are used to alter the
mobility of the driving fluid and/or to reduce the capillary forces [133].
Two major forces are acting on phase entrapment in the pore structure: viscous
forces and capillary forces. It was shown that the capillary number must be increased
from 10–6 to 10–2 for mobilization of entrapped oil ganglia [78]. The enhanced oil
recovery mechanisms are divided into two types. In the first type the ratio of viscous
to capillary forces is altered and in the second type the fluid phase volume and the
10 Capillary Phase Trapping 441

Fig. 10.20 Flowsheet of EOR techniques

preferential wettability of the porous medium, i.e. the residual phase’s configuration,
are altered. According to the capillary number, in the oil–water system by reducing
interfacial tension to 10–2 –10–4 dynes/cm, more than 90% recovery efficiency is
gained [134]. So the reduction in interfacial tension leads to mobilization of oil
ganglia. Generally, Pore geometry, interfacial tension, and wettability were found
to be the most important parameters which control ganglion mobilization and affect
capillary trapping.
The flowsheet of EOR techniques is shown in Fig. 10.20.

10.6.1 Gas Injection to Remove Trapped Phase

Displacement of reservoir liquids by gas occurs in various geological applications


such as enhanced oil recovery [135], carbon storage and sequestration [136], and envi-
ronmental remediation [137]. The efficacy of these subsurface operations depends on
the degree of capillary trapping which is lead to the distribution of each fluid phase in
the porous media [123]. The transport behaviour of two immiscible phases depends
on fluid properties such as viscosity ratio, interfacial tension and the textural proper-
ties of a porous medium such as pore topology and surface roughness [123]. Mehmani
et al. quantified the effects of surface roughness on viscous fingering morphology and
capillary trapping during gas injection employing microfluidics experiments [123].
Gas injection investigations concentrating on fluid properties have typically focused
on the effects of viscosity ratio and capillary number [138–140]. Knowledge of the
permeability, pore geometry, and initial fluid saturation is essential to properly eval-
uate the trapped phase. The displacement mechanism during gas injection could be
described with the drainage process. Snap-off was first characterized by Roof [22]
in a drainage process using equilibrium assumptions, i.e. pressure gradients within
phases are ignored. The higher the pore aspect ratio, the more likely it is for snap-off
to occur. At low capillary numbers, however, potentially due to the low pore aspect
ratio, snap-off has typically been neglected in the conceptual models for drainage
442 F. Kazemi et al.

[65, 141, 142]. Recent snap-off observations of gas injection into core samples using
X-ray micro tomography have shown that snap-off could occur even at low pore
aspect ratios due to non-local pressure/velocity oscillations and demonstrated a
cooperative pore mechanism for snap-off [143].
During gas injection, large quantities of oil can be by-passed. The amount of by-
passing depends on the heterogeneity, wettability of the medium and enrichment of
the injection fluids. Subsequent mass transfer from the by-passed regions can recover
a significant fraction of this oil. Four mechanisms take part in this mass transfer:
dispersion/diffusion-drive [144], gravity-drive [145], pressure-drive and capillarity-
drive [146].
It was investigated that water saturation can affect by-passing and mass transfer,
and decrease mass transfer between oil and gas. It is important to note that the
capillary forces become more dominant as enrichment decreases [147].
Lean Gas Injection
Dry lean gas injection is an effective method for trapped condensate removal in gas
condensate reservoir. If sufficient bottom-hole injection pressures can be attained
(generally above 35,000 kPa) vaporizing miscibility can be obtained between conven-
tional dry natural gas or nitrogen and many light volatile condensates. In this proce-
dure, since the injected gas is lean and contains no heavy ends, it can vaporize a
considerable amount of the trapped liquid hydrocarbons [148].
Rich Gas Injection
If injection pressures are too low to facilitate vaporizing miscibility with lean gas,
richer gases such as liquid CO2 or ethane have been used to obtain a similar effect
at much lower pressures. For even lower pressures, richer solvents such as propane
and butane have also been used to remove trapped condensate.
It is important to ensure that the rich gas is compatible with the condensate and will
not cause the undesirable precipitation of asphaltenes or other potentially damaging
solids [148].
Water Alternating Gas Injection
Water alternating gas (WAG) process is one of the oil recovery methods that involve
three-phase fluid flow. WAG injection is originally proposed as a method to improve
the sweep efficiency of gas by using water to control the mobility ratio and to stabilize
the front [149, 150].
Multiple displacements during WAG injection is an important mechanism for
the trapping and remobilization of gas and oil, especially in media with a low
coordination number and poor connectivity of wetting and spreading layers [151].
Piri and Blunt studied the multiple displacement process in double drainage where
gas displaces trapped oil that displaces water [152]. In WAG flooding, double and
multiple displacement processes involving trapped gas will also be important, studied
by Suicmez et al. [153]. It is important to notice that, experimental three-phase flow
measurements are extremely difficult to perform [154–156].
10 Capillary Phase Trapping 443

In water-wet system, the gas is the most non-wetting phase and also the most
mobile phase. So the gas relative permeability controls the efficiency of the process.
In WAG injection, gas invades the largest pores and throats. During wetting phase
injection (water or oil), gas is displaced from the smaller pores. The saturation of the
gas phase may not be affected significantly, but this is lead to disconnection of contin-
uous gas clusters due to by-passing trapping, and therefore reducing the gas phase
relative permeability. By continuing gas/water injection, the gas relative permeability
and recovery efficiency decreased [157–159]. In mixed-wet or oil-wet reservoir, the
gas is not necessarily the most non-wetting phase, so it can be the intermediate-wet
phase in a strongly oil-wet medium and the displacement mechanisms and also pore
occupancies are expected to be different [160, 161].
Reduction of Oil Viscosity
The injected gas can dissolve in the oil and makes the oil volume expand. So the inter-
molecular distance increase, the intermolecular force decrease, and the oil viscosity
decrease. Therefore the oil seepage resistance in the formation is reduced, resulting
in facilitating the oil production in the formation and enhanced oil recovery [162].
Reduction of Interfacial Tension
The CO2 injection leads to extraction and gasification of the light components of the
oil. So the oil–water interfacial tension be decreased and caused enhance oil recovery.
Furthermore, if the pressure of the injection is greater than the minimum miscible
pressure, CO2 can be miscible with the oil and lead to a reduction of oil–water
interfacial tension to zero and enhancing the oil recovery [162].
Changing Morphology
CO2 can react with formation water to generate H2 CO3 , which can react with the
rock. So the improvement of the permeability of the oil layer is happened [162].
Enhanced Oil-Water Density Ratio
The light components of oil in the formation can be extracted by injected CO2 , and
the oil density is increased. So the difference of the density between oil and water is
reduced, and in the subsequent waterflooding the enhanced oil recovery is achieved
[162].
Enhanced Oil-Water Fluidity Ratio
The fluidity of oil will be increased by dissolving CO2 in oil since the viscosity of
oil decreased. On the other hand, when CO2 is dissolved in the water, the viscosity
of formation water will be increased, and lead to decreasing the fluidity of water.
Finally, the reduction of water–oil fluidity ratio enlarges the swept volume of the
subsequent waterflooding [162].
Oil Expansion during Gas Injection
After CO2 injection, it is dissolved in the oil of the formation under the conditions of
its temperature and pressure and expands the volume of the oil. Therefore more oil
444 F. Kazemi et al.

enables to be produced from the formation and lead to enhancing the good production
and the recovery [162].

10.7 Capillary Trapping Models

Residual saturation data have traditionally been compared to several empirical trap-
ping models. All trapping models have been proposed based on experimental data
and different rock types, wetting, and non-wetting fluids, and different experimental
conditions.

10.7.1 Land Model (1968)

The most widely used model was proposed by Land [163] base on a two-phase exper-
imental study on a water-wet sandstone cores [163]. This model contains a constant,
C, called the Land trapping coefficient, which is computed from the endpoints of
water flood curves and depends on the type of rock and fluid properties. When
C = 0, all the non-wetting phase is trapped, and when C = ∞, no trapping occurs
[164]. It is defined as:

Snwi
Snwr = (10.9)
1 + C Snwi

And
 
1 1
C= max
− max
(10.10)
Snwr Snw

where S nwi and S nwr are the non-wetting phase saturation after primary drainage and
after secondary imbibition, respectively. Snwr max is the maximum of Snwr , and Snw max
is the maximum of trapped oil saturation. A weakness of Land’s model is that it is
only based on the endpoints. Also, it has been shown that some initial-residual (IR)
trapping experiments cannot be predicted by the Land trapping model and need to
rely on other empirical trapping models.

10.7.2 Aissaoui Model (1983)

Aissaoui proposed a bi-linear trapping model that matched the Suzanne et al. [165]
experimental data set. Recently at Imperial College London [30, 91, 166] it was
shown that the best match to experimental data from unconsolidated sand packs is
10 Capillary Phase Trapping 445

given by Aissaioui’s law. It features two linear components. Initially, Sgr increases
linearly with Sgi before plateauing at Sgr max [24, 165].
Aissaoui showed that Snwr initially increased linearly with Snwi and then became
constant at the maximum non-wetting phase saturation, Snwr max . This equation
contains Snwc , which is the critical non-wetting phase saturation corresponding to
the maximum trapped non-wetting saturation. It is defined as:
max
Snwr
I f Snwi < Snwc , then Snwr = Snwi ; else Snwr = Snwr
max
(10.11)
Snwc

10.7.3 Ma and Youngren Model (1994)

One adaptation of Land’s relationship was proposed by Ma and Youngren (1994).


They found that the IR curve of the oil-wet system sharply levelled off at high initial
non-wetting phase saturations (gas phase saturation). This led to the introduction of
two empirically curve fitting parameters in their equation, a and b, where a = C and
b = 1 in Land’s original correlation [41]:

Snwi
Snwr = (10.12)
1 + aSnwi
b

As mentioned, a shortcoming of the Land model is that it is based upon the


endpoint only, and there is no scope to tune the shape of the IR curve [30]. One model
that went some way toward tackling this limitation is Ma and Youngren model [30].

10.7.4 Jerauld Model (1997)

Another correlation that adapted Land’s model was proposed by Jerauld [27, 167].
This model can predict the amount of the non-wetting phase saturation in mixed-wet
rock and contains a parameter that depends on the non-wetting phase saturation (gas
or oil) that accounts for the “plateau” observed in the IR curves for mixed-wet rocks.
The trapped nonwetting-phase saturation is given by:

Snwi
Snwr =   (10.13)
1+ 1
max
Snwr
− 1 (Snwi )a

for gas:

1 + bSnwr
max
a= (10.14)
1 − Snwr
max
446 F. Kazemi et al.

for oil:
1
a= (10.15)
1 − Snwr
max

Jerauld introduced a second tuning parameter b in addition to the Land coefficient.


If this parameter is set to zero, Jerauld’s model reduces to the Land trapping model.
When this parameter is equal to one, the trapping curve has a zero slope at Snwi =
1. Although Jerauld argued that the IR curves should not have a negative slope, his
model allows for such behaviour if b ≥ 1.

10.7.5 Kleppe et al. Model (1997)

Kleppe et al. proposed a linear relationship between initial and residual saturations
to match their experimental data based on an artificial core [168]. They found that
Land’s trapping model couldn’t predict their experimental data. The following linear
relationship was proposed to match the data set [168]:

Snwi max
Snwr = max Snwr (10.16)
Snwi

This model agrees with Aissaoui model.

10.7.6 Spiteri et al. Model (2008)

Spiteri et al. showed that the IR curves may not be fitted to a linear curve and
may have a parabolic trend and proposed a quadratic correlation based on pore-
network modelling for mixed-wet and oil-wet conditions. They established a model
differed from Land’s approach and used a second-order polynomial to match their
experimental data which has two fitting parameters were based on the contact angle
[36, 120]:

Snwr = αSnwi − β Snwi


2
(10.17)

where α and β are fitting constants [120].


Besides, Spiteri et al. showed that Jerauld’s model could match the experimental
data of water-wet and intermediate-wet media, but that it was not good for oil-wet
systems [120, 169].
10 Capillary Phase Trapping 447

10.7.7 Iglauer et al. Model (2009)

Iglauer et al. statistically analysed oil-brine and air-brine laboratory data for to find
a relationship between Sor and Soi and Sgr and Sgi in unconsolidated systems [24]. In
the least square approach, bi-linear relationships were derived [170, 171] as bellow:
For the oil-brine experimental data, it was found:

0.2535Soi i f Soi ≤ 0.403
Sor = (10.18)
0.0891 + 0.0324Soi i f Soi ≥ 0.403

And for the air-brine experimental data:



0.7165Sgi i f Sgi ≤ 0.194
Sgr = (10.19)
0.1384 + 0.00114Sgi i f Sgi ≥ 0.194

As mentioned before, in the context of CCS the capillary trapping capacity (Ctrap )
is of great importance since it states how much CO2 can be stored securely per unit
rock volume:
It was found a correlation for their oil-brine experimental data:

0.0938Snwi i f Snwi ≤ 0.403
Ctrap = (10.20)
0.03296 + 0.0120Snwi i f Snwi ≥ 0.403

And for air-brine experimental data:



0.2651Snwi i f Snwi ≤ 0.194
Ctrap = (10.21)
0.05135 + 0.000411Snwi i f Snwi ≥ 0.194

Iglauer et al. also proposed residual saturation as a function of porosity and


trapping capacity as a function of porosity using least square fits [24].

10.7.8 Evaluation of Best Trapping Model

Different trapping models that represent the relationship between initial and residual
saturations were evaluated in both oil–water and gas–water systems [29]. The accu-
racy of each model was evaluated by different experimental data. The model param-
eters were fitted by curve fitting, and suitable models for different types of rock and
fluids was proposed. As shown in Tables 10.1 and 10.2 the Spiteri et al. and Ma
and Youngren models could predict residual saturations better than other trapping
models. The maximum R2’ s of the Spiteri et al. and Ma and Youngren models were
448

Table 10.1 Fitting parameters and errors of different trapping models in oil–water (brine) systems and water-wet media
Land Kleppe et al. Aissaoui Ma and Youngren Jerauld Spiteri et al. Refs.
C R2 Snwi max Snwr max R2 Snwr max Swc R2 a b R2 a R2 α β R2
2.307 0.303 0.602 0.252 −6.191 0.252 0.602 −6.191 2.931 1.471 1.000 1.337 0.670 0.898 0.798 0.989 [37]
1.632 0.999 0.489 0.272 0.787 0.272 0.489 0.787 1.562 0.934 0.999 1.374 0.916 0.858 0.624 0.996 [37]
2.628 0.785 0.533 0.220 −1.467 0.222 0.533 −1.467 3.133 1.321 0.955 1.285 0.712 0.863 0.834 0.993 [37]
1.828 0.115 0.771 0.320 0.636 0.320 0.642 0.258 1.577 −0.147 0.718 1.471 −0.220 0.324 0.075 0.340 [32]
0.741 −0.011 0.914 0.486 0.925 0.486 0.914 0.925 2.764 0.662 0.240 1.330 −2.982 0.449 0.192 0.237 [34]
0.856 0.949 0.859 0.495 0.805 0.495 0.859 0.805 0.979 1.421 0.959 1.980 0.933 0.972 0.482 0.955 [36]
0.465 0.527 0.748 0.555 0.538 0.555 0.743 0.527 0.938 1.605 0.850 2.247 0.657 1.018 0.519 0.852 [36]
0.835 0.678 0.935 0.525 0.788 0.525 0.858 0.849 0.861 0.406 0.817 2.105 0.230 0.686 0.154 0.824 [36]
7.008 0.788 0.841 0.122 0.692 0.122 0.646 0.897 6.441 0.591 0.939 1.139 0.662 0.318 0.223 0.971 [40]
5.938 0.845 0.796 0.139 0.735 0.139 0.737 0.819 5.270 0.586 0.934 1.161 0.759 0.370 0.264 0.972 [30]
0.907 0.972 0.892 0.493 0.801 0.493 0.892 0.801 1.005 1.595 0.988 1.972 0.979 0.990 0.501 0.989 [39]
0.964 0.909 0.914 0.486 0.925 0.486 0.914 0.925 0.952 0.456 0.950 1.945 0.782 0.697 0.195 0.951 [9]
F. Kazemi et al.
Table 10.2 Fitting parameters and errors of different trapping models in gas–water (brine) systems and water-wet media
Land Kleppe et al Aissaoui Ma and Youngren Jerauld Spiteri et al. Refs.
C R2 Snwi max Snwr max R2 Snwr max Swc R2 a b R2 a R2 α β R2
1.533 0.588 0.855 0.370 0.379 0.370 0.754 0.498 1.842 0.952 0.759 0.831 0.752 0.751 0.450 0.764 [92]
1.605 0.046 0.809 0.352 −2.677 0.352 0.674 −1.112 2.699 2.339 0.744 1.666 0.576 1 0.750 0.608 [41]
10 Capillary Phase Trapping

2.108 0.849 0.799 0.298 0.280 0.298 0.536 0.765 3.019 1.632 0.901 1.304 0.883 0.916 0.767 0.911 [27]
1.821 0.416 0.808 0.327 −1.933 0.327 0.808 −1.933 2.495 1.727 0.763 1.404 0.672 0.983 0.775 0.651 [90]
1.467 −0.050 0.715 0.349 −2.759 0.349 0.672 −1.995 3.009 2.716 0.955 1.938 0.819 1 0.702 0.676 [90]
1.219 −2.026 0.701 0.378 −2.469 0.378 0.586 −2.628 4.433 2.505 −0.003 0.947 −0.828 1.166 1.178 0.041 [90]
1.950 0.408 0.678 0.292 −1.087 0.292 0.678 −1.087 3.239 1.713 0.704 1.298 0.612 0.956 0.868 0.620 [90]
1.662 0.194 0.657 0.314 −1.032 0.314 0.503 −0.378 3.533 1.884 0.687 1.243 0.508 1.045 1.031 0.672 [90]
5.374 0.564 0.801 0.151 −2.101 0.151 0.640 −1.406 6.843 1.403 0.824 1.185 0.701 0.670 0.668 0.500 [40]
1.560 0.886 0.853 0.366 0.958 0.366 0.853 0.958 1.361 0.205 0.965 0.629 0.931 0.544 0.139 0.972 [39]
2.821 −0.095 0.494 0.206 −5.643 0.206 0.363 −0.416 5.612 1.763 0.577 1.382 0.423 0.972 1.234 0.640 [93]
1.825 0.922 0.724 0.312 0.729 0.312 0.651 0.905 1.757 0.864 0.927 1.314 0.894 0.781 0.501 0.959 [35]
3.983 0.650 0.754 0.188 −0.739 0.188 0.598 0.926 3.404 0.711 0.738 1.118 0.545 0.481 0.298 0.838 [35]
2.845 0.824 0.728 0.237 0.810 0.237 0.724 0.818 2.238 0.428 0.908 1.051 0.783 0.532 0.275 0.939 [35]
0.882 0.935 0.681 0.425 0.746 0.425 0.671 0.787 0.818 0.789 0.948 1.527 0.879 0.857 0.351 0.952 [33]
1.523 0.451 0.583 0.309 0.173 0.309 0.583 0.173 1.945 1.090 0.814 1.264 0.798 0.829 0.605 0.824 [33]
0.861 0.849 0.772 0.464 0.257 0.464 0.757 0.345 1.168 1.818 0.920 1.799 0.920 1.077 0.652 0.921 [33]
449
450 F. Kazemi et al.

1 and 0.99 for oil/brine systems, and 0.95 and 0.97 for gas-brine fluid pairs, respec-
tively, regardless of the type of porous media. These two models also accurately fitted
the Initial-Residual (IR) experimental data of the gas–oil system.

10.8 Capillary Trapping Laboratory Methods

There are several laboratory methods for measuring the amount of trapped phase
and hence achieving trapping curve, which is the relationship between the initial
non-wetting phase saturation (Snwi ) and the residual non-wetting phase saturation
(Snwr ). These methods include centrifuge, unsteady-state, steady-state, and porous
plate discussed in the following subsection.
To measure the initial and residual saturation, volumetric and material balance
method, can be used as shown in the following equations.

Voi − VDV
V olumetric Method : Soi = ,
PV
Soi × P V − Vw f
Sor = (10.22)
PV
Mdr y − Mdragnage + (ρbrine P V )
Material Balance Method : Soi = ,
P V (ρbrine − ρoil )
Mdr y − Mw f + (ρbrine P V )
Sor = (10.23)
P V (ρbrine − ρoil )

where
Soi is the initial oil saturation.
Sor is the residual oil saturation.
Voi is the volume of displaced water with oil during primary drainage.
VDV is the dead volume of cell.
P V is the pore volume.
Vw f is the volume of displaced oil during water flooding.
Mdr y is the mass of the dry core.
Mdragnage is the mass of the core after drainage.
Mw f is the mass of the core after water flooding.
ρbrine is the density of brine.
ρoil is the density of oil.
10 Capillary Phase Trapping 451

Centrifuge
arm

Rock

Graduated
Glass tube

Fig. 10.21 Schematic of the centrifuge capillary trapping experiment a Drainage process.
b Imbibition process

10.8.1 Centrifuge Method

In this method, both the drainage and imbibition capillary pressure curves can be
gained. For drainage, a core sample filled with the wetting phase is placed in a cell
surrounded by the non-wetting phase. Then the cell centrifuged at several speeds
starting from low to high. At each speed, the amount of fluid displaced during the
process is measured. The schematic of this method is illustrated in Fig. 10.21.
The process continues until no further production is seen at high speeds [172].
For the imbibition curve, the cell will be reversed compared to drainage since the oil
phase is usually less dense than water.

10.8.2 Unsteady State (USS) Method

In an unsteady state method, as shown in Fig. 10.22, the core is saturated with one
fluid (wetting or non-wetting fluid), and then the other fluid is injected to displace
the first fluid. During unsteady state displacement, the flow rates of both fluids and
pressure drop are recorded. Using this method is cheap and relatively quick. In this
method the relative permeability is not measured directly so requires interpretation
of the results [173]. However, if this method is used strictly to achieve the trapping
curve, the method becomes simpler and less time-consuming.
452 F. Kazemi et al.

∆P

Oil
Brine

Brine
L

Fig. 10.22 Schematic of an unsteady state capillary trapping coreflood experiment

∆P

Oil Oil

Brine Brine
L

Fig. 10.23 Schematic of a steady-state capillary trapping coreflood experiment

10.8.3 Steady State (SS) Method

In the steady-state measurement, both fluids are injected simultaneously to the rock
as illustrated in Fig. 10.23. This method is primarily used to measure the relative
permeability [172]. To find the relative permeability, the flow rate and pressure drop
should be measured. However, the trapping curve can be achieved as a by-product
and doesn’t require additional stages.

10.8.4 Porous Plate (PP) Method

Porous plate (PP), which is a semi-permeable water-wet ceramic disc, retain the oil
in the core and let the water pass through the outlet (Fig. 10.24). The advantages of
using this method are the ability to establish various initial saturations and eliminate
the capillary end effect, a saturation gradient from the inlet to the outlet of the core,
ability to establish low initial saturations compared to the unsteady state technique,
and also to establish a homogeneous distribution of saturation inside the core. The
disadvantage of this method is that it is time-consuming [5, 9, 35].
10 Capillary Phase Trapping 453

Brine Brine

(a)

Porous plate
Oil Brine

(b)

Oil
Brine Brine

(c)

Fig. 10.24 Schematic of a porous plate capillary trapping coreflood experiment

Appendix: Derivation of Velocity Ratio Expression

The purpose of this Appendix is to drive the velocity ratio expression in the pore
doublet model, which determine the interface position (Fig. 10.25).
The magnitude of capillary pressure affect the velocity of the interface in both
capillary tube. The static capillary pressure function (Pc equation) is defined as
Eq. 10.23.

2σ cos θ
Pc = (10.24)
R

where σ is the interfacial tension between invading and invaded fluid, θ is the
advancing contact angle and R is the radius of the capillary.

Wetting phase R1

R2 Non-wetting phase

ΔP

Fig. 10.25 Schematic of the pore doublet structure


454 F. Kazemi et al.

Furthermore, the volumetric flow rate in either path is given by Hagen-Poiseuille


(H-P) equation:

π R 4 P
q= (10.25)
8μ L

where R is the radius of the capillary, L is the length of the flow path of the doublet,
assumed to be the same in both branches and μ is the viscosity, assuming that
both fluids have the same viscosity. The magnitude of viscose pressure drop can be
calculated using Eq. 10.24, if the flow rate was known. Equation 10.24, implies the
assumptions of cylindrical capillaries of circular cross-section and of fully developed
and steady laminar flows that none of which is likely to be met in reality.
The total volumetric flow rate through the doublet, if the water is supplied at the
volume flow rate, qw , to the doublet can be written as:
π  4 
qw = q1 + q2 = R1 P1 + R24 P2 (10.26)
8μL

where ql is volume flow rates in capillary 1 and q2 is volume flow rates in capillary 2.
The flow direction from left to right are taken with a positive sign and in reverse direc-
tion with a negative sign. It is assumed that the interfaces advance simultaneously in
both capillaries of pore doublet.
The pressure drop across each capillary is the sum of the capillary pressure and
the viscous pressure drop. Since the two capillaries are parallel, the net pressure
difference between the two branch points must be the same when taken over either
of the two capillaries 1 and 2.

P1 − Pc1 = P2 − Pc2 (10.27)

And so

Pc = Pc2 − Pc1 = P2 − P1 (10.28)

In Eq. 10.26, the capillary pressure is positive for imbibition process and negative
for drainage one.
Equation (10.27) can be rearranged as:

P1 = P2 − Pc (10.29)

Substitute Eqs. (10.28) into (10.25):


π  4 
qw = R1 P2 − R14 Pc + R24 P2 (10.30)
8μL
10 Capillary Phase Trapping 455

So Eq. (10.29) can be rearranged as:

π P2  4  π R14
qw = R1 + R24 − Pc (10.31)
8μL 8μL

Then, P2 can be found using Eq. (10.30).


 
π R14 Pc 8μL
P2 = qw + ∗  4  (10.32)
8μL π R1 + R24

Equation (10.25) it can be rearranged as:

π R24
q1 = qw − q2 = qw − P2 (10.33)
8μL

Then Eqs. (10.32) together with Eq. (10.31) results in:


 
π R24 π R14 Pc 8μL
q1 = qw − ∗ qw + ∗  4  (10.34)
8μL 8μL π R1 + R24

Equation 10.33 can be simplified to:


⎛ ⎞   ⎛ ⎞
R24 π R14 R24 Pc R14 π R14 R24 Pc
q1 = qw −   qw − ⎝   ⎠ = qw − ⎝   ⎠
R14 + R24 8μL R14 + R24 R14 + R24 8μL R14 + R24
(10.35)
 
Dividing this equation by R1 4 and multiplying by R14 + R24 :
   
q1 R14 + R24 π R24 Pc
= qw − (10.36)
R14 8μL

So the equation of q1 and q2 are as follows:


   
π R24 Pc π R24 σ cos θ
qw − 8μL
qw − 4μL
1
R2
− 1
R1
q1 =  4 =  4 (10.37)
1+ R2
R1
1+ R2
R1
 4  4   4  
π R2 Pc π R24 σ cos θ
R2
R1
qw + 8μL
R2
R1
q w + 4μL
1
R2
− 1
R1
q2 =  4 =  4 (10.38)
1 + RR21 1 + RR21

Then expressions of the velocities, v1 and v2, will be obtained by dividing each
π D2 π D2
flow rate by its respective cross-sectional area ( 4 1 and 4 2 ):
456 F. Kazemi et al.

    
π R24 Pc q R24
q ∗ 12 − ∗ 12 2 − ∗ (σ cos θ) ∗ 1 1
R2 − R1
q1 π R1 8μL π R1 π R1 4μL R12
v1 = =  4 =  4
π R12 1+ R
R2 R
1 + R2
1 1
(10.39)
      
R2 4 π R24 Pc R22 R22
1
R1 q ∗ π R 2 + ∗ 12 ∗ πq + 4μL ∗ (σ cos θ) ∗ 1 1
R2 − R1
q2 8μL π R2 R4
v2 = = 2
 4 = 1
 4
π R22 R2
1+ R
R
1 + R2
1 1
(10.40)

And
So dividing these two velocity to find the ratio of velocity:
  
R22 R22
v2 R14
∗ q
π
+ 4μL
∗ (σ cos θ ) ∗ R12 − R11
=    (10.41)
v1 q

R24
∗ (σ cos θ) ∗ R12 − R11
π R12 4μL R12

Dividing this expression by R22 σ cos θ and multiplying by 4μL R1


   
4μLq
v2 π R13 σ cos θ
+ R1
R2
−1 4Ncap + 1
β
−1
=  =   (10.42)
v1 4μLq

R22
∗ 1
− 1 4Ncap
− β2 ∗ 1
−1
π R1 R22 σ cos θ R1 R2 R1 β2 β

Finally, Eq. 10.19 can be further simplified to:


 
v2 4Ncap + 1
β
−1
=   (10.43)
v1 4Ncap
− β2 ∗ 1
−1
β2 β

In which
μLq
Ncap = (10.44)
π R13 σ cos θ

And
R2
β= (10.45)
R1
10 Capillary Phase Trapping 457

References

1. Tzimas E, Georgakaki A, Cortes CG, Peteves SD. Enhanced oil recovery using carbon dioxide
in the European energy system. Rep EUR. 2005;21895.
2. Harper EJ. Optimization of capillary trapping of CO2 sequestration in saline aquifers; 2012.
3. White CM, Strazisar BR, Granite EJ, Hoffman JS, Pennline HW. Separation and capture of
CO2 from large stationary sources and sequestration in geological formations—coalbeds and
deep saline aquifers. J Air Waste Manage Assoc. 2003;53:645–715.
4. Riazi M, Sohrabi M, Bernstone C, Jamiolahmady M, Ireland S. Visualisation of mechanisms
involved in CO2 injection and storage in hydrocarbon reservoir sand water-bearing aquifers.
Chem Eng Res Des. 2011;89:1827–40.
5. Alyafei N. Capillary trapping and oil recovery in altered-wettability carbonate rock; 2015.
6. Herring AL, Andersson L, Schlüter S, Sheppard A, Wildenschild D. Efficiently engineering
pore-scale processes: the role of force dominance and topology during nonwetting phase
trapping in porous media. Adv Water Resour. 2015;79:91–102.
7. Davis BBJ, Wood WD. Maximizing economic return by minimizing or preventing aqueous
phase trapping during completion and stimulation operations. In: SPE annual technical
conference and exhibition. Society of Petroleum Engineers; 2004.
8. Raeini AQ, Bijeljic B, Blunt MJ. Modelling capillary trapping using finite-volume simulation
of two-phase flow directly on micro-CT images. Adv Water Resour. 2015;83:102–10.
9. Tanino Y, Blunt MJ. Laboratory investigation of capillary trapping under mixed-wet
conditions. Water Resour Res. 2013;49:4311–9.
10. Lenormand R, Touboul E, Zarcone C. Numerical models and experiments on immiscible
displacements in porous media. J Fluid Mech. 1988;189:165–87.
11. Morrow NR, Songkran B. Effect of viscous and buoyancy forces on nonwetting phase trapping
in porous media. In: Surface phenomena in enhanced oil recovery. Springer; 1981. p. 387–411.
12. Iglauer S, Wu Y, Shuler P, Tang Y, Goddard WA III. New surfactant classes for enhanced oil
recovery and their tertiary oil recovery potential. J Pet Sci Eng. 2010;71:23–9.
13. Shah DO. Improved oil recovery by surfactant and polymer flooding. Elsevier; 2012.
14. Teletzke GF, Patel PD, Chen A. Methodology for miscible gas injection EOR screening.
In: SPE/DOE symposium on improved oil recovery conference in Asia Pacific. Society of
Petroleum Engineers; 2005.
15. Li S, Genys M, Wang K, Torsæter O. Experimental study of wettability alteration during
nanofluid enhanced oil recovery process and its effect on oil recovery. In: SPE reservoir
characterisation and simulation conference and exhibition. Society of Petroleum Engineers;
2015
16. Al-Anazi HA, Xiao J, Al-Eidan AA, Buhidma IM, Ahmed MS, Al-Faifi M, et al. Gas
productivity enhancement by wettability alteration of gas-condensate reservoirs. In: European
formation damage conference. Society of Petroleum Engineers; 2007
17. Wang Y, Xu H, Yu W, Bai B, Song X, Zhang J. Surfactant induced reservoir wettability
alteration: recent theoretical and experimental advances in enhanced oil recovery. Pet Sci.
2011;8:463–76.
18. Kumar A, Noh MH, Ozah RC, Pope GA, Bryant SL, Sepehrnoori K, et al. Reservoir simulation
of CO2 storage in aquifers. SPE J. 2005;10:336–48.
19. Kwelle SO. Experimental studies on resistance to fluid displacement in single pores; 2017.
20. Herring AL, Gilby FJ, Li Z, McClure JE, Turner M, Veldkamp JP, et al. Observations of
nonwetting phase snap-off during drainage. Adv Water Resour. 2018;121:32–43.
21. Pentland CH, Iglauer S, Gharbi O, Okada K, Suekane T. The influence of pore space geometry
on the entrapment of carbon dioxide by capillary forces. In: SPE Asia Pacific oil and gas
conference and exhibition. Society of Petroleum Engineers; 2012.
22. Roof JG. Snap-off of oil droplets in water-wet pores. Soc Pet Eng J. 1970;10:85–90.
23. Blunt MJ, Scher H. Pore-level modeling of wetting. Phys Rev E. 1995;52:6387.
458 F. Kazemi et al.

24. Wulling W, Pentland CH, Al Mansoori SK, Blunt MJ. Capillary trapping capacity of rocks
and sandpacks. In: EUROPEC/EAGE Conference and Exhibition. Society of Petroleum
Engineers; 2009
25. Tanino Y, Blunt MJ. Capillary trapping in sandstones and carbonates: dependence on pore
structure. Water Resour Res 2012;48.
26. Aissaoui A. Etude théorique et expérimentale de l’hystérésis des pressions capillaires et des
perméabilités relatives en vue du stockage souterrain de gaz. Ec Des Mines Paris, Paris; 1983.
27. Jerauld GR. Prudhoe Bay gas/oil relative permeability. SPE Reserv Eng. 1997;12:66–73.
28. Delclaud J. Laboratory measurements of the residual gas saturation. In: Second society of
core analysts European core analysis symposium; 1991. p. 431–51.
29. Kazemi F, Azin R, Osfouri S. Evaluation of phase trapping models in gas-condensate systems
in an unconsolidated sand pack. J Pet Sci Eng. 2020;195:107848.
30. Pentland CH, Itsekiri E, Al-Mansoori S, Iglauer S, Bijeljic B, Blunt MJ. Measurement of
nonwetting-phase trapping in sandpacks. SPE J. 2010;15:274–81.
31. Alyafei N, Blunt MJ. The effect of wettability on capillary trapping in carbonates. Adv Water
Resour. 2016;90:36–50.
32. Masalmeh SK. The effect of wettability on saturation functions and impact on carbonate reser-
voirs in the Middle East. In: Abu Dhabi international petroleum exhibition and conference.
Society of Petroleum Engineers; 2002.
33. Akbarabadi M, Piri M. Relative permeability hysteresis and capillary trapping characteristics
of supercritical CO2 /brine systems: an experimental study at reservoir conditions. Adv Water
Resour. 2013;52:190–206.
34. Skauge A, Sørvik A, Vik B, Spildo K. Effect of wettability on oil recovery from carbonate
material representing different pore classes. In: Proceedings of the international symposium
of the society of core analysts; 2006. p. 12–6.
35. El-Maghraby RM, Blunt MJ. Residual CO2 trapping in Indiana limestone. Environ Sci
Technol. 2012;47:227–33.
36. Pentland CH, Tanino Y, Iglauer S, Blunt MJ. Capillary trapping in water-wet sandstones:
coreflooding experiments and pore-network modeling. In: SPE conference. SPE International;
2010.
37. Pickell JJ, Swanson BF, Hickman WB. Application of air-mercury and oil-air capillary
pressure data in the study of pore structure and fluid distribution. Soc Pet Eng J. 1966;6:55–61.
38. Pentland CH. Measurements of non-wetting phase trapping in porous media; 2011.
39. Pentland CH, El-Maghraby R, Iglauer S, Blunt MJ. Measurements of the capillary trapping
of super-critical carbon dioxide in Berea sandstone. Geophys Res Lett. 2011;38.
40. Al Mansoori SK, Itsekiri E, Iglauer S, Pentland CH, Bijeljic B, Blunt MJ. Measurements
of non-wetting phase trapping applied to carbon dioxide storage. Int J Greenh Gas Control.
2010;4:283–8.
41. Ma TD, Youngren GK. Performance of immiscible water-alternating-gas (IWAG) injection at
Kuparuk River Unit, North Slope, Alaska. In: SPE annual technical conference and exhibition.
Society of Petroleum Engineers; 1994.
42. Craig FF. The reservoir engineering aspects of waterflooding, vol. 3. HL Doherty Memorial
Fund of AIME New York; 1971.
43. Corey AT. Mechanics of immiscible fluids in porous media. Water Resources Publication;
1994.
44. Sohrabi M, Danesh A, Tehrani DH, Jamiolahmady M. Microscopic mechanisms of oil
recovery by near-miscible gas injection. Transp Porous Media. 2008;72:351–67.
45. Buckley JS. Effective wettability of minerals exposed to crude oil. Curr Opin Colloid Interface
Sci. 2001;6:191–6.
46. Buckley JS, Bousseau C, Liu Y. Wetting alteration by brine and crude oil: from contact angles
to cores. SPE J. 1996;1:341–50.
47. Anderson WG. Wettability literature survey-part 1: rock/oil/brine interactions and the effects
of core handling on wettability. J Pet Technol. 1986;38:1,121–125,144.
10 Capillary Phase Trapping 459

48. Anderson WG. Wettability literature survey part 5: the effects of wettability on relative
permeability. J Pet Technol. 1987;39:1, 451–453, 468.
49. Gharbi O, Blunt MJ. The impact of wettability and connectivity on relative permeability in
carbonates: a pore network modeling analysis. Water Resour Res. 2012;48.
50. Land CS. Comparison of calculated with experimental imbibition relative permeability. Soc
Pet Eng J. 1971;11:419–25.
51. Longino BL. Residual non-wetting phase retention and mobilization in rough-walled
fractures. Queen’s University at Kingston; 1999.
52. Wilson JL. Laboratory investigation of residual liquid organics from spills, leaks, and
the disposal of hazardous wastes in groundwater. Robert S. Kerr Environmental Research
Laboratory, Office of Research and …; 1990.
53. Eikevåg TK. Experimental study of residual gas saturation using both spontaneous and forced
imbibition method, where IsoparL is the wetting phase. Institutt for petroleumsteknologi og
anvendt geofysikk; 2012.
54. Van Duijn CJ, Molenaar J, De Neef MJ. The effect of capillary forces on immiscible two-phase
flow in heterogeneous porous media. Transp Porous Media. 1995;21:71–93.
55. Chaouche M, Rakotomalala N, Salin D, Xu B, Yortsos YC. Capillary effects in drainage
in heterogeneous porous media: continuum modelling, experiments and pore network
simulations. Chem Eng Sci. 1994;49:2447–66.
56. Van Batenburg DW, Bruining J, Bakken H, Palmgren CTS. The effect of capillary forces in
heterogeneous “flow-units”: a streamline approach. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1991
57. Or D. Scaling of capillary, gravity and viscous forces affecting flow morphology in unsaturated
porous media. Adv Water Resour. 2008;31:1129–36.
58. Zhang C, Oostrom M, Wietsma TW, Grate JW, Warner MG. Influence of viscous and capil-
lary forces on immiscible fluid displacement: pore-scale experimental study in a water-wet
micromodel demonstrating viscous and capillary fingering. Energy Fuels. 2011;25:3493–505.
59. Bear J. Dynamics of fluids in porous media. Courier Corporation; 2013.
60. Suekane T, Nobuso T, Hirai S, Kiyota M. Geological storage of carbon dioxide by residual
gas and solubility trapping. Int J Greenh Gas Control. 2008;2:58–64.
61. Shook M, Li D, Lake LW. Scaling immiscible flow through permeable media by inspectional
analysis. IN SITU-NEW YORK. 1992;16:311.
62. Heiss VI, Neuweiler I, Ochs S, Färber A. Experimental investigation on front morphology
for two-phase flow in heterogeneous porous media. Water Resour Res. 2011;47.
63. Morrow NR, Chatzis I, Taber JJ. Entrapment and mobilization of residual oil in bead packs.
SPE Reserv Eng. 1988;3:927–34.
64. Willhite GP. Waterflooding, vol. 3. Richardson, Texas Textbook Series SPE 1986.
65. Lenormand R, Zarcone C, Sarr A. Mechanisms of the displacement of one fluid by another
in a network of capillary ducts. J Fluid Mech. 1983;135:337–53.
66. Herring AL. An investigation into the pore-scale mechanisms of capillary trapping: application
to geologic CO2 sequestration; 2014.
67. Chatzis I, Morrow NR. Correlation of capillary number relationships for sandstone. Soc Pet
Eng J. 1984;24:555–62.
68. Cense AW, Berg S. The viscous-capillary paradox in 2-phase flow in porous media. In: Inter-
national symposium of the society of core analysts held in Noordwijk. Netherlands; 2009.
p. 27–30.
69. Ding M, Kantzas A. Capillary number correlations for gas-liquid systems. In: Canadian
international pet conference. Petroleum Society of Canada; 2004.
70. Shen P, Zhu B, Li X-B, Wu Y-S. An experimental study of the influence of interfacial tension
on water–oil two-phase relative permeability. Transp Porous Media. 2010;85:505–20.
71. Chatzis I, Kuntamukkula MS, Morrow NR. Effect of capillary number on the microstructure
of residual oil in strongly water-wet sandstones. SPE Reserv Eng. 1988;3:902–12.
72. Bona N, Garofoli L, Radaelli F, Zanaboni C, Bendotti A, Mezzapesa D. Trapped gas saturation
measurements: new perspectives. In: SPE annual technical conference and exhibition. Society
of Petroleum Engineers; 2014.
460 F. Kazemi et al.

73. Lake LW. Enhanced oil recovery; 1989.


74. Dake JMK. Essentials of engineering hydraulics; 1983.
75. Coninck H, Loos M, Metz B, Davidson O, Meyer L. IPCC special report on carbon dioxide
capture and storage. Intergov Panel Clim Chang; 2005.
76. Chatzis I, Dullien FAL. Dynamic immiscible displacement mechanisms in pore doublets:
theory versus experiment. J Colloid Interface Sci. 1983;91:199–222.
77. Taber JJ. Dynamic and static forces required to remove a discontinuous oil phase from porous
media containing both oil and water. Soc Pet Eng J. 1969;9:3–12.
78. Melrose JC. Role of capillary forces in detennining microscopic displacement efficiency for
oil recovery by waterflooding. J Can Pet Technol. 1974;13.
79. Foster WR. A low-tension waterflooding process. J Pet Technol. 1973;25:205–10.
80. Flett M, Gurton R, Taggart I. The function of gas-water relative permeability hysteresis in
the sequestration of carbon dioxide in saline formations. In: SPE Asia Pacific oil and gas
conference and exhibition. Society of Petroleum Engineers; 2004.
81. Juanes R, Spiteri EJ, Orr FM, Blunt MJ. Impact of relative permeability hysteresis on
geological CO2 storage. Water Resour Res. 2006;42.
82. Juanes R, MacMinn C. Upscaling of capillary trapping under gravity override: application
to CO2 sequestration in aquifers. In: SPE symposium on improved oil recovery. Society of
Petroleum Engineers; 2008.
83. Qi R, LaForce TC, Blunt MJ. Design of carbon dioxide storage in aquifers. Int J Greenh Gas
Control. 2009;3:195–205.
84. Hesse M, Tchelepi HA, Orr FM. Scaling analysis of the migration of CO2 in saline aquifers.
In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2006.
85. Obi EI, Blunt MJ. Streamline-based simulation of carbon dioxide storage in a north sea aquifer.
Water Resour Res. 2006;42.
86. Saadatpoor E, Bryant SL, Sepehrnoori K. Effect of heterogeneous capillary pressure on
buoyancy-driven CO2 migration. In: SPE symposium on improved oil recovery. Society of
Petroleum Engineers; 2008.
87. Ide ST, Jessen K, Orr FM Jr. Storage of CO2 in saline aquifers: Effects of gravity, viscous, and
capillary forces on amount and timing of trapping. Int J Greenh Gas Control. 2007;1:481–91.
88. Hawkes CD, McLellan PJ, Bachu S. Geomechanical factors affecting geological storage of
CO2 in depleted oil and gas reservoirs. J Can Pet Technol. 2005;44.
89. Lackner KS. A guide to CO2 sequestration. Science (80-) 2003;300:1677–8.
90. Kralik JG, Manak LJ, Jerauld GR, Spence AP. Effect of trapped gas on relative permeability
and residual oil saturation in an oil-wet sandstone. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 2000.
91. Al Mansoori SK, Iglauer S, Pentland CH, Blunt MJ. Three-phase measurements of oil and
gas trapping in sand packs. Adv Water Resour. 2009;32:1535–42.
92. Crowell DC, Dean GW, Loomis AG. Efficiency of gas displacement from a water-drive
reservoir, vol. 6735. US Department of the Interior, Bureau of Mines; 1966.
93. Krevor SCM, Pini R, Zuo L, Benson SM. Relative permeability and trapping of CO2 and
water in sandstone rocks at reservoir conditions. Water Resour Res 2012;48.
94. Alafnan S, Aljawad M, Alismail F, Almajed A. Enhanced recovery from gas condensate
reservoirs through renewable energy sources. Energy Fuels. 2019;33:10115–22.
95. Martinez RGF. Altering wettability in gas condensate sandstone reservoirs for gas mobillity
improvement. Texas A & M University; 2012.
96. Fan L, Harris B, Jamaluddin A. Understanding gas-condensate reservoirs. Oilf Rev. 2005:14–
27. https://doi.org/10.4043/25710-MS.
97. Fulcher RA Jr, Ertekin T, Stahl CD. Effect of capillary number and its constituents on two-
phase relative permeability curves. J Pet Technol. 1985;37:249–60.
98. Azin R, Sedaghati H, Fatehi R, Osfouri S, Sakhaei Z. Production assessment of low production
rate of well in a supergiant gas condensate reservoir: application of an integrated strategy. J
Pet Explor Prod Technol. 2019;9:543–60.
10 Capillary Phase Trapping 461

99. Sakhaei Z, Mohamadi-Baghmolaei M, Azin R, Osfouri S. Study of production enhance-


ment through wettability alteration in a super-giant gas-condensate reservoir. J Mol Liq.
2017;233:64–74.
100. Mohamadi-Baghmolaei M, Azin R, Osfouri S, Zendehboudi S. Evaluation of mass transfer
coefficient for gas condensates in porous systems: experimental and modeling. Fuel.
2019;255:115507.
101. Ahmadi R, Osfouri S, Azin R. Wettability alteration of carbonate oil reservoir surface using
biocompatible nanoparticles. Mater Res Express. 2018;6:25033.
102. Mohamadi-Baghmolaei M, Azin R, Osfouri S, Zendehboudi S. Experimental and modeling
investigation of non-equilibrium condensate vaporization in porous systems: effective
determination of mass transfer coefficient. Fuel. 2020;262:116011.
103. Luo K, Li S, Zheng X, Chen G, Dai Z, Liu N. Experimental investigation into revaporization
of retrograde condensate by lean gas injection. In: SPE Asia Pacific oil and gas conference
and exhibition. Society of Petroleum Engineers; 2001
104. Khadar RH, Aminshahidy B, Hashemi A, Ghadami N. Application of gas injection and
recycling to enhance condensate recovery. Pet Sci Technol. 2013;31:1057–65.
105. Settari A, Bachman RC, Hovem KA, Paulsen SG. Productivity of fractured gas condensate
wells-a case study of the Smorbukk field. SPE Reserv Eng. 1996;11:236–44.
106. Yuan C, Zhang Z, Liu K. Assessment of the recovery and front contrast of CO2 EOR and
sequestration in a new gas condensate reservoir by compositional simulation and seismic
modeling. Fuel. 2015;142:81–6.
107. Kossack CA, Opdal ST. Recovery of condensate from a heterogeneous reservoir by the injec-
tion of a slug of methane followed by nitrogen. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 1988.
108. Sanger PJ, Hagoort J. Recovery of gas-condensate by nitrogen injection compared with
methane injection. SPE J. 1998;3:26–33.
109. Jamaluddin AKM, Ye S, Thomas J, D’Cruz D, Nighswander J. Experimental and theoretical
assessment of using propane to remediate liquid buildup in condensate reservoirs. In: SPE
annual technical conference and exhibition. Society of Petroleum Engineers; 2001.
110. Antoci JC, Briggiler NJ, Chadwick JA. Crosslinked methanol: analysis of a successful expe-
rience in fracturing gas wells. In: SPE Latin American and Caribbean petroleum engineering
conference. Society of Petroleum Engineers; 2001.
111. Barnum RS, Brinkman FP, Richardson TW, Spillette AG. Gas condensate reservoir behaviour:
productivity and recovery reduction due to condensation. In: SPE annual technical conference
and exhibition. Society of Petroleum Engineers; 1995.
112. Du L, Walker JG, Pope GA, Sharma MM, Wang P. Use of solvents to improve the productivity
of gas condensate wells. In: SPE annual technical conference and exhibition. Society of
Petroleum Engineers; 2000.
113. Ahmadi R, Farmani Z, Osfouri S, Azin R. Condensate blockage remediation in a gas reservoir
through wettability alteration using natural CaCO3 nanoparticles. Colloids Surf Physicochem
Eng Asp. 2019;579:123702.
114. Naghizadeh A, Azin R, Osfouri S, Fatehi R. Wettability alteration of calcite and dolomite
carbonates using silica nanoparticles coated with fluorine groups. J Pet Sci Eng. 2020:106915.
115. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate
cores by alumina-nanofluid in different base fluids and temperature. J Sustain Energy Eng.
2018;6:84–98.
116. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate
rocks from strongly liquid-wetting to strongly gas-wetting by fluorine-doped silica coated by
fluorosilane. J Dispers Sci Technol. 2018;39:767–76.
117. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Synthesis of fluorine-doped silica-
coating by fluorosilane nanofluid to ultrahydrophobic and ultraoleophobic surface. Mater Res
Express. 2017;4:105010.
118. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Stability of alumina nanofluid in
water/methanol base fluid in the presence of different salts. J Nanofluids. 2018;7:235–45.
462 F. Kazemi et al.

119. Sakhaei Z, Azin R, Naghizadeh A, Osfouri S, Saboori R, Vahdani H. Application of fluori-


nated nanofluid for production enhancement of a carbonate gas-condensate reservoir through
wettability alteration. Mater Res Express. 2018;5:35008.
120. Spiteri EJ, Juanes R, Blunt MJ, Orr FM. A new model of trapping and relative permeability
hysteresis for all wettability characteristics. SPE J. 2008;13:277–88.
121. Hu R, Wan J, Kim Y, Tokunaga TK. Wettability impact on supercritical CO2 capillary trapping:
pore-scale visualization and quantification. Water Resour Res. 2017;53:6377–94.
122. Princ T, Fideles HMR, Koestel J, Snehota M. The impact of capillary trapping of air on satiated
hydraulic conductivity of sands interpreted by X-ray microtomography. Water. 2020;12:445.
123. Mehmani A, Kelly S, Torres-Verdín C, Balhoff M. Residual oil saturation following gas injec-
tion in sandstones: Microfluidic quantification of the impact of pore-scale surface roughness.
Fuel. 2019;251:147–61.
124. Oh SG, Slattery JC. Interfacial tension required for significant displacement of residual oil.
Soc Pet Eng J. 1979;19:83–96.
125. Chatzis I, Morrow NR, Lim HT. Magnitude and detailed structure of residual oil saturation.
Soc Pet Eng J. 1983;23:311–26.
126. Wardlaw NC. The effects of geometry, wettability, viscosity and interfacial tension on trapping
in single pore-throat pairs. J Can Pet Technol. 1982;21.
127. Nguyen VH, Sheppard AP, Knackstedt MA, Pinczewski WV. The effect of displacement rate
on imbibition relative permeability and residual saturation. J Pet Sci Eng. 2006;52:54–70.
128. Rose W, Witherspoon PA. Studies of waterflood performance. Division of the Ill. State Geol.
Survey; 1956.
129. Manceau J-C, Hatzignatiou DG, De Lary L, Jensen NB, Réveillère A. Mitigation and reme-
diation technologies and practices in case of undesired migration of CO2 from a geological
storage unit—current status. Int J Greenh Gas Control. 2014;22:272–90.
130. Kong X, Ohadi M. Applications of micro and nano technologies in the oil and gas industry-
overview of the recent progress. In: Abu Dhabi international petroleum exhibition and
conference. Society of Petroleum Engineers; 2010.
131. Company BP. BP statistical review of world energy June 2005. British Petroleum Educational
Service; 2005.
132. Green DW, Willhite GP. Enhanced oil recovery, vol. 6. Henry L. Doherty Memorial Fund of
AIME, Society of Petroleum Engineers Richardson, TX; 1998.
133. Shah DO. Surface phenomena in enhanced oil recovery. Springer; 1981.
134. Stegemeier GL. Mechanisms of entrapment and mobilization of oil in porous media. Improv
Oil Recover Surfactant Polym Flood. 1997:55–91.
135. Bath PGH. Status report on miscible/immiscible gas flooding. J Pet Sci Eng. 1989;2:103–17.
136. DePaolo DJ, Cole DR. Geochemistry of geologic carbon sequestration: an overview. Rev
Mineral Geochemistry. 2013;77:1–14.
137. Marley MC, Hazebrouck DJ, Walsh MT. The application of in situ air sparging as an innovative
soils and ground water remediation technology. Groundw Monit Remediat. 1992;12:137–45.
138. Lake LW, Carey GF, Pope GA, Sepehrnoori K. Isothermal, multiphase, multicomponent fluid
flow in permeable media. In Situ (United States). 1984;8.
139. Al-Gharbi MS, Blunt MJ. 2D dynamic pore-scale network model of imbibition. Dev Water
Sci. 2004;55: 71–82 (Elsevier).
140. Doorwar S, Mohanty KK. Fingering function for unstable immiscible flows. In: SPE reservoir
simulation symposium. Society of Petroleum Engineers; 2015.
141. Joekar-Niasar V, Doster F, Armstrong RT, Wildenschild D, Celia MA. Trapping and hysteresis
in two-phase flow in porous media: a pore-network study. Water Resour Res. 2013;49:4244–
56.
142. Valvatne P. Predictive porescale modeling of single and multiphase flow. Transp Porous Media.
2004;55:71–89.
143. Andrew M, Menke H, Blunt MJ, Bijeljic B. The imaging of dynamic multiphase fluid
flow using synchrotron-based X-ray microtomography at reservoir conditions. Transp Porous
Media. 2015;110:1–24.
10 Capillary Phase Trapping 463

144. Hara SK, Christman PG. Investigation of a cyclic countercurrent light-oil/CO2 immiscible
process. SPE Adv Technol Ser. 1993;1:159–65.
145. Firoozabadi A, Tan JCT. Miscible displacement in fractured porous media: part II-analysis.
In: SPE/DOE improved oil recovery symposium. Society of Petroleum Engineers; 1994.
146. Pande KK, Orr FM Jr. Analytical computation of breakthrough recovery for CO2 floods in
layered reservoirs. Pap SPE. 1990;20177:22–5.
147. Wylie P, Mohanty KK. Effect of water saturation on oil recovery by near-miscible gas injection.
SPE Reserv Eng. 1997;12:264–8.
148. Bennion DB, Thomas FB, Bietz RF, Bennion DW. Remediation of water and hydrocarbon
phase trapping problems in low permeability gas reservoirs. J Can Pet Technol. 1999;38.
149. Caudle BH, Dyes AB. Improving miscible displacement by gas-water injection. Trans AIME.
1958;213:281–3.
150. Christensen JR, Stenby EH, Skauge A. Review of WAG field experience. In: International
petroleum conference and exhibition of Mexico. Society of Petroleum Engineers; 1998.
151. Van Dijke MIJ, Sorbie KS. Pore-scale modelling of three-phase flow in mixed-wet porous
media: multiple displacement chains. J Pet Sci Eng. 2003;39:201–16.
152. Piri M, Blunt MJ. Three-dimensional mixed-wet random pore-scale network modeling of
two-and three-phase flow in porous media. II. Results. Phys Rev E. n.d.;71:26302.
153. Suicmez VS, Piri M, Blunt MJ. Pore-scale simulation of water alternate gas injection. Transp
Porous Media. 2007;66:259–86.
154. Oak MJ. Three-phase relative permeability of intermediate-wet Berea sandstone. In: SPE
annual technical conference and exhibition. Society of Petroleum Engineers; 1991.
155. Blunt MJ. An empirical model for three-phase relative permeability. SPE J. 2000;5:435–45.
156. Baker LE. Three-phase relative permeability correlations. In: SPE enhanced oil recovery
symposium. Society of Petroleum Engineers; 1988.
157. Egermann P, Vizika O, Dallet L, Requin C, Sonier F. Hysteresis in three-phase flow: experi-
ments, modeling and reservoir simulations. In: SPE European petroleum conference. Society
of Petroleum Engineers; 2000.
158. Element DJ, Masters JHK, Sargent NC, Jayasekera AJ, Goodyear SG. Assessment of three-
phase relative permeability models using laboratory hysteresis data. In: SPE/DOE symposium
on improved oil recovery conference in Asia Pacific. Society of Petroleum Engineers; 2003.
159. Feng Q, Di L, Tang G, Chen Z, Wang X, Zou J. A visual micro-model study: The mechanism
of water alternative gas displacement in porous media. In: SPE/DOE symposium on improved
oil recovery. Society of Petroleum Engineers; 2004.
160. DiCarlo DA, Akshay S, Blunt MJ. Three-phase relative permeability of water-wet, oil-wet,
and mixed-wet sandpacks. SPE J. 2000;5:82–91.
161. Dicarlo DA, Sahni A, Blunt MJ. The effect of wettability on three-phase relative permeability.
Transp Porous Media. 2000;39:347–66.
162. Wu Y, John JC, Yongle H. The three sisters: acid gas injection, carbon capture and
sequestration, and enhanced oil recovery; 2019. p. 7.
163. Land CS. Calculation of imbibition relative permeability for two- and three-phase flow from
rock properties. Soc Pet Eng J. 1968;8:149–56. https://doi.org/10.2118/1942-PA.
164. Spiteri EJ, Juanes R. Impact of relative permeability hysteresis on the numerical simulation
of WAG injection. J Pet Sci Eng. 2006;50:115–39.
165. Suzanne K, Hamon G, Billiotte J, Trocme V. Experimental relationships between residual gas
saturation and initial gas saturation in heterogeneous sandstone reservoirs. In: SPE annual
technical conference and exhibition. Society of Petroleum Engineers; 2003.
166. Gittins P, Iglauer S, Pentland CH, Al-Mansoori S, Al-Sayari S, Bijeljic B, et al. Nonwetting
phase residual saturation in sand packs. J Porous Media. 2010;13.
167. Jerauld GR. General three-phase relative permeability model for Prudhoe Bay. SPE Reserv
Eng. 1997;12:255–63.
168. Kleppe J, Delaplace P, Lenormand R, Hamon G, Chaput E. Representation of capillary pres-
sure hysteresis in reservoir simulation. In: SPE annual technical conference and exhibition.
Society of Petroleum Engineers; 1997.
464 F. Kazemi et al.

169. Spiteri E, Juanes R, Blunt MJ, Orr FM. Relative-permeability hysteresis: trapping models
and application to geological CO2 sequestration. In: SPE annual technical conference and
exhibition. Society of Petroleum Engineers; 2005.
170. Breiman L. Probability. SIAM, Philadelphia. Math Rev 93d. 1992;60001.
171. Lawson CL, Hanson RJ. Solving least squares problems, vol. 15. Siam; 1995.
172. Tiab D, Donaldson EC. Petrophysics: theory and practice of measuring reservoir rock and
fluid transport properties. Gulf Professional Publishing; 2015.
173. Dullien FAL. Porous media: fluid transport and pore structure. Academic Press; 2012.

You might also like