You are on page 1of 256

Petroleum Production Technology

采油技术
Changhong Gao
高长虹
Responsible Editor: Qunxia Wan

图书在版编目(CIP)数据
采油技术 = Petroleum Production Technology: 英文版 / 高长虹著. —北京:科学出版社, 2018. 4
Ⅰ. ①采… Ⅱ. ①高… Ⅲ. ①石油开采–基本知识–英文 Ⅳ. ①TE35
中国版本图书馆 CIP 数据核字(2018)第 072947 号

Published by Science Press


16 Donghuangchenggen North Street
Beijing 100717, P. R. China
Printed in Beijing

Copyright© 2017 by Science Press


ISBN 978-7-03-057210-3

All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording or otherwise, without the prior written permission of the copyright owner.
About the Author
Changhong Gao received his degrees in petroleum
engineering from China University of Petroleum, University of
Tulsa, and Curtin University. He worked for Sinopec as a field
engineer and research engineer. He also taught various petroleum
engineering courses at United Arab Emirates University,
University of Aberdeen, and Petroleum Institute of Abu Dhabi.
His teaching and research interests include production system
design, enhanced oil recovery, and drilling mud. He has
published more than 30 journal papers and presented a few papers at SPE and APPEA
conferences. In 2013, he was elected Taishan Scholar by Shandong province
government, as well as Yellow River Delta Scholar by Dongying city government.
The author can be contacted at: 237184689@qq.com
Preface
After some years of experiences in both university and industry, the author
realizes there is a demand for a book that covers the complete petroleum production
chain.
This book covers the most important aspects of petroleum production technology.
Chapter 1 and chapter 2 present properties of oil, gas, and reservoir rock. Chapter 3
introduces mechanisms of primary recovery, secondary recovery, and enhanced oil
recovery. Chapter 4 presents equations that govern fluid flow in reservoir. Chapter 5
introduces onshore and offshore facilities for drilling and production. Chapter 6
presents options for lower and upper completions. Chapter 7 introduces well
productivity index and inflow performance relationship. Chapter 8 presents
single-phase flow and multiphase flow in wells, as well as nodal analysis technique for
production optimization. Chapter 9 introduces artificial lift methods, including sucker
rod pump, gas lift, and electrical submersible pump etc. Chapter 10 presents techniques
for well stimulation, namely acidizing and fracturing. Chapter 11 and chapter 12
introduce common production problems and remediation techniques, such as sand,
scale, wax, hydrate and asphaltene. Chapter 13 presents separation and treatment for oil,
gas, and water.
After reading this book, the readers gain valuable knowledge in modern
production technology in the most concise and straightforward manner, without
struggling through tedious and excessive descriptions. A few illustrative examples are
included to make theories easy to understand. This book serves as a textbook for
students studying petroleum engineering, as well as a reference book for engineers in
petroleum industry.
The Author is grateful for the financial support from Shandong province
government through the prestigious Taishan Scholar program. Dongying city
government also provided financial support through Yellow River Delta Scholar
program.
rof. Yongho
·ii· Petroleum Production Technology
Contents ·iii·

Contents
Preface
Chapter 1 Properties of Oil and Gas........................................................................ 1
1.1 Phase behavior ............................................................................................... 2
1.2 Properties of natural gas ................................................................................ 5
1.2.1 Gas equation of state ............................................................................... 6
1.2.2 Gas compressibility ................................................................................ 9
1.2.3 Gas formation volume factor ................................................................. 10
1.2.4 Gas viscosity ........................................................................................ 10
1.3 Properties of crude oil ................................................................................. 12
1.3.1 Oil density ........................................................................................... 12
1.3.2 Solution gas oil ratio ............................................................................. 13
1.3.3 Oil formation volume factor .................................................................. 14
1.3.4 Oil viscosity ......................................................................................... 16
1.4Fluid sampling ............................................................................................. 17
Chapter 2 Properties of Reservoir Rock................................................................ 20
2.1 Basic petroleum geology ............................................................................. 20
2.2 Porosity ........................................................................................................ 22
2.3 Absolute permeability ................................................................................. 24
2.4 Fluid Saturation ........................................................................................... 29
2.5 Surface tension and interfacial tension ........................................................ 30
2.6 Wettability ................................................................................................... 30
2.7 Capillary action ........................................................................................... 33
2.8 Capillary pressure ........................................................................................ 34
2.9 Relative permeability................................................................................... 39
2.10 Electrical conductivity ............................................................................... 43
2.11 Rock sampling ........................................................................................... 44
Chapter 3 Reservoir Recovery Mechanisms ......................................................... 48
3.1 Field life cycle ............................................................................................. 48
3.2 Primary recovery mechanisms..................................................................... 50
3.2.1 Rock and liquid expansion drive ............................................................ 50
3.2.2 Solution gas drive ................................................................................. 50
3.2.3 Gas cap drive ....................................................................................... 51
·iv· Petroleum Production Technology

3.2.4 Aquifer drive ........................................................................................ 53


3.2.5 Combination drive ................................................................................ 54
3.3 Secondary recovery methods ....................................................................... 55
3.4 Enhanced oil recovery ................................................................................. 62
3.4.1 Thermal methods .................................................................................. 62
3.4.2 Miscible gas flood ................................................................................ 63
3.4.3 Chemical methods ................................................................................ 66
Chapter 4 Fluid Flow in Reservoir ........................................................................ 71
4.1 Basic concepts ............................................................................................. 71
4.2 Basic diffusivity equation ............................................................................ 72
4.3 Linear flow under steady state ..................................................................... 75
4.3.1 Steady-state linear flow of incompressible fluid ....................................... 75
4.3.2 Steady-state linear flow of slightly compressible fluid .............................. 76
4.3.3 Steady-state linear flow of compressible gases ........................................ 77
4.4 Radial flow under steady state ..................................................................... 80
4.4.1 Steady-state radial flow of incompressible fluid under.............................. 80
4.4.2 Steady-state radial flow of compressible gases ........................................ 84
4.5Fluid flow under pseudo-steady state .......................................................... 86
Chapter 5 Drilling and Production Facilities ........................................................ 90
5.1 Drilling rigs ................................................................................................. 90
5.2 Drilling systems ........................................................................................... 93
5.2.1 Rotary drive and top drive ..................................................................... 93
5.2.2 Drill string ........................................................................................... 95
5.2.3 Mud circulation .................................................................................... 97
5.2.4 Blowout preventer ................................................................................ 97
5.2.5 Casing program .................................................................................... 98
5.3Offshore production structures .................................................................. 100
Chapter 6 Well Completion .................................................................................. 108
6.1 Lower completion ...................................................................................... 108
6.1.1 Open-hole completion ......................................................................... 108
6.1.2 Cased-hole completion .........................................................................112
6.1.3 Choice of lower completion ..................................................................112
6.1.4 Cementing a cased well ........................................................................113
6.1.5 Perforating a cased well .......................................................................118
Contents ·v·

6.2 Upper completion ...................................................................................... 122


6.2.1 Options for upper completion .............................................................. 122
6.2.2 Christmas tree and tubing hanger ......................................................... 124
6.2.3 Safety valve ....................................................................................... 127
6.2.4 Production packer ............................................................................... 129
6.2.5 Smart completion ............................................................................... 130
Chapter 7 Well Inflow Performance .................................................................... 132
7.1 Well productivity index ............................................................................. 132
7.2 Vogel IPR................................................................................................... 133
7.3 Fetkovich IPR ............................................................................................ 140
7.4Productivity of Horizontal Wells ............................................................... 141
Chapter 8 Fluid Flow in Wells and Nodal Analysis ............................................ 147
8.1 Single-phase flow in pipes......................................................................... 147
8.2 Multiphase flow concept ........................................................................... 150
8.3 Multiphase flow models ............................................................................ 155
8.4 Tubing performance relationship ............................................................... 161
8.5 Nodal analysis for production system........................................................ 163
Chapter 9 Artificial Lift Methods ........................................................................ 167
9.1 Sucker rod pump........................................................................................ 167
9.2 Gas lift ....................................................................................................... 172
9.2.1 Gas lift system.................................................................................... 172
9.2.2 Well unloading and kick-off ................................................................ 174
9.3 Electrical submersible pump ..................................................................... 179
9.4 Other artificial lift methods ....................................................................... 182
9.5Discussions ................................................................................................ 186
Chapter 10 Acidizing and Fracturing .................................................................. 187
10.1 Formation damage and skin factor .......................................................... 187
10.2 Fracturing ................................................................................................ 190
10.2.1 Fracture pressure .............................................................................. 190
10.2.2 Fracturing fluid and proppants ........................................................... 192
10.2.3 Models of fracture geometry .............................................................. 193
10.3Matrix acidizing....................................................................................... 195
Chapter 11 Sand Control ...................................................................................... 199
11.1 Grain size distribution.............................................................................. 199
·vi· Petroleum Production Technology

11.2 Gravel pack .............................................................................................. 202


11.3 Screens ..................................................................................................... 203
11.3.1 Wire-wrapped screens ....................................................................... 203
11.3.2 Premium screens ............................................................................... 205
11.3.3 Pre-packed screens ............................................................................ 206
11.3.4 Expandable screens ........................................................................... 207
Chapter 12 Flow Assurance ...................................................................................211
12.1 Scale .........................................................................................................211
12.2 Waxes ...................................................................................................... 213
12.3 Asphaltenes.............................................................................................. 216
12.4 Hydrates................................................................................................... 218
12.4.1 Types of hydrates.............................................................................. 219
12.4.2 Hydrate prevention ........................................................................... 222
12.4.3 Removal of hydrate plug ................................................................... 224
Chapter 13 Separation and Treatment ................................................................ 226
13.1 Separator configurations .......................................................................... 226
13.2 Theories of separation ............................................................................. 231
13.2.1 Terminal velocity .............................................................................. 231
13.2.2 Gas capacity constraint ...................................................................... 232
13.2.3 Retention time requirement ................................................................ 233
13.2.4 Oil pad thickness .............................................................................. 234
13.2.5 Water phase thickness ....................................................................... 235
13.2.6 Slenderness ratio ............................................................................... 236
13.3 Separator sizing ....................................................................................... 236
13.4 Gas treatment ........................................................................................... 237
13.4.1 Gas dehydration ................................................................................ 237
13.4.2 Gas compression ............................................................................... 239
13.4.3 Downstream gas processing ............................................................... 240
13.5 Oil dehydration ........................................................................................ 241
13.6 Water treatment........................................................................................ 241
Main References ....................................................................................................... 243
Appendix A................................................................................................................ 245
Chapter 1 Properties of Oil and Gas
It is necessary for engineers in oil and gas industry to be equipped with knowledge in
properties of petroleum fluids. This chapter presents the most important concepts related
to properties of natural gas and crude oil. This chapter covers the definitions of key
parameters, as well as the most widely used methods to predict oil and gas properties.
Originally, hydrocarbon and water are sealed in reservoir rocks. Before the
reservoir comes on stream, the fluids in the reservoir are maintained under initial
reservoir pressure and temperature. After the production begins, fluids flow through
reservoir and wellbore to reach surface. During this process, the pressure and temperature
decline, and the properties of oil and gas also change with changing conditions.
Crude oil is a mixture of light to heavy hydrocarbons, plus some sulphur, oxygen,
nitrogen and metallic components. According to different structures, hydrocarbons are
classified as paraffin (alkane), olefin (alkene), naphthene and aromatics. Some
examples of alkane series are given in Fig. 1-1. Alkane are open chain molecules with
saturated bonds, They have the generalized formula CnH2n+2. Each carbon atom has
four open bonds, thus can join with four other atoms.

Fig. 1-1 Some examples of alkanes


·2· Petroleum Production Technology

Under standard conditions, methane, ethane, propane and butane are in gaseous
state. The density of alkanes increases with increasing carbon number. Under standard
conditions, C5 to C17 remain in liquid state, while the C18 and higher compounds exist
in wax-like solid state. Alkanes from C1 to C40 typically appear in crude oil. They
contribute up to 20% of crude oil by volume.
Alkenes have at least one carbon-carbon double bond. The double bond makes
alkenes more reactive than alkanes. Naphthenes are ring structures with saturated
bonds, while aromatics are rings with double bonds, as shown in Fig. 1-2.

Fig. 1-2 Examples of ring structures

1.1 Phase behavior

Phase behavior describes the phase in which a mass of fluid exists at given conditions
of pressure, volume and temperature (PVT). The simplest way to start to understand
this relationship is by considering a single component, for example, water, under only
two variables: Pressure and temperature.
Fig. 1-3 shows the phase behavior in the solid, liquid and vapor states. Starting
with the liquid phase at point A, as the temperature is increased, the boiling point is
reached, at which the water boils and turns to steam (gas). Starting from the situation
of the gaseous phase at point B, if the temperature is reduced, the dew point curve is
reached, and the component changes from the gaseous phase to the liquid phase. For a
single component, the boiling point curve and the dew point curve are coincident, and
are known as the vapor pressure curve. The phase boundary between the liquid and
solid phases is the melting point curve. At the triple point, all three phases coexist. At
critical point, it is impossible to distinguish between liquid and gas phases. The
Chapter 1 Properties of Oil and Gas ·3·

compressed gas exists in liquid state.

Fig. 1-3 Phase behavior of a pure component

Reservoir fluids may exist as vapor or liquid, even as solid phase. The change in
phases can be directly mapped on a pressure-temperature (P-T ) diagram. While a
reservoir produces fluids, its pressure depletes, and its temperature remains relatively
constant.
A typical P-T diagram is shown in Fig. 1-4. The phase envelope is formed by
connecting bubble point curve and dew point curve. The bubble point curve separates
liquid phase from the two-phase region. Two phases coexist inside the envelope. The
two curves converge at critical point. The pressure and temperature at critical point are
named critical pressure and critical temperature. Cricondentherm refers to the
maximum temperature beyond which liquid cannot form regardless of pressure.
Cricondenbar refers to the maximum pressure beyond which gas cannot form
regardless of temperature. Different types of oil and gas reservoirs are defined based on
their P-T diagrams.
·4· Petroleum Production Technology

Fig. 1-4 A typical P-T diagram for crude oil(Unit:%)

If the reservoir temperature is lower than the critical temperature of the reservoir
fluid, the reservoir is referred to as an oil reservoir. Oil reservoirs can be further
classified into three types as follows. ① Saturated reservoir: If the initial reservoir
pressure is equal to the bubble point pressure of reservoir fluid, the reservoir is referred
to as saturated reservoir. ② Undersaturated reservoir: If the initial reservoir pressure
is higher than the bubble point pressure of reservoir fluid, the reservoir is referred to as
undersaturated reservoir. ③ Gas cap reservoir: If the initial reservoir pressure is
below the bubble point pressure of reservoir fluid, the reservoir is referred to as gas cap
reservoir. In this scenario, a gas zone lies above the oil zone. The gas that exists in the
gas cap is named free gas, while that dissolved in the oil phase is named solution gas.
Crude oils can be classified as volatile oil or black oil. A volatile oil contains large
fractions of light and intermediate component which vaporizes easily. Large amount of
gas is liberated with a small drop in pressure below bubble point. Black oils contain a
lower fraction of light components, thus requiring a more significant pressure drop
below bubble point for gas to liberate from oil phase.
If the reservoir temperature is above the critical temperature of hydrocarbon in the
reservoir, the reservoir is classified as a gas reservoir. Gas reservoirs can be further
classified as three types. ① Dry gas reservoir: For dry gas reservoir, the produced gas
always exists as vapor phase, the only liquid produced is water. ② Wet gas reservoir:
For this type of reservoir, its temperature is above the cricondentherm, Hence the
Chapter 1 Properties of Oil and Gas ·5·

original hydrocarbon is in vapor state. However, while the gas flows to surface, its
pressure and temperature reduce and a liquid phase begins to condense. This type of
reservoir is classified as wet gas reservoir. ③ Retrograde condensate reservoir: If the
reservoir temperature lies between the critical temperature and the cricondentherm of
the reservoir hydrocarbon system, the reservoir is classified as retrograde condensate
reservoir.
Retrograde condensate demonstrates very unique thermodynamic behavior, which
is illustrated in Fig. 1-5. At point 1 in Fig. 1-5, the hydrocarbon exists as vapor phase.
As pressure decline, it expands in volume. While pressure further declines into the
two-phase region, it begins to condense rather than expand further. The condensate
volume reaches maximum at point 3. Afterwards, the condensate begins to vaporize
while pressure declines further (point 4).

Fig. 1-5 P-T diagram for retrograde condensate

1.2 Properties of natural gas

Natural gas is a naturally occurring hydrocarbon gas mixture consisting primarily of


methane, but commonly includes varying amounts of other higher alkanes, and sometimes a
small percentage of carbon dioxide, nitrogen, hydrogen sulfide, or helium. Natural gas is an
important source of energy for heating, cooking, and electricity generation. It is also a
major feedstock for production of plastics and fertilizers.
·6· Petroleum Production Technology

1.2.1 Gas equation of state


When a gas is stored at ambient pressure and temperature, it behaves like an ideal gas.
When pressure and temperature increase, its behavior deviates from ideal gas. The
ideal gas law is given below:
PV = n RT (1-1)

where,
P = pressure, psi;
V = volume, ft3;
T = absolute temperature, ºR;
n = number of moles of gas, mol;
R = universal gas constant (10.73 for the above units).
It is clear that
n=m/M (1-2)

where,
m = the mass of gas, g;
M = the gas molecular weight, g/mol.
Therefore,
m
PV = RT (1-3)
M
And the density of gas can be expressed as:
m PM
ρg = = (1-4)
V RT
Natural gas is a mixture of components. The apparent molecular weight of the
mixture depends on the molecular weight of each component and its fraction:
n
M a = ∑ yi M i (1-5)
i =1

where,
Ma = apparent molecular weight of a gas mixture;
Mi = molecular weight of a component in the mixture;
yi = mole fraction of a component in the mixture.
The specific gravity of gas is defined as the ratio of gas density to air density.
Chapter 1 Properties of Oil and Gas ·7·

ρg
γg = (1-6)
ρair

Incorporating the density, the specific gravity of gas can be expressed as:
Ma
γg = (1-7)
29
For gases under high pressure, the ideal gas law leads to high errors. The ideal gas
law must be adjusted to suit real gases under high pressure and high temperature.
Therefore, a parameter named gas compressibility factor (or gas Z factor) is
introduced. The real gas law is expressed as:
PV = ZnRT (1-8)
where,
R = universal gas constant.
psi① ⋅ ft 3
The value of gas constant is 10.73 with the field units. The
lb − mol ⋅ °R
determination of gas Z factor follows the steps below.①
(1) Calculate pseudo-critical properties:
n
Ppc = ∑ yi Pci (1-9)
i =1
n
Tpc = ∑ yi Tci (1-10)
i =1

If the composition is unknown, the pseudo-critical properties can be estimated by


the following equations:
Ppc = 677 + 15γg − 37.5γg2

Tpc = 168 + 325γg − 12.5γg2

(2) Calculate pseudo-reduced pressure and temperature:


P
Ppr = (1-11)
Ppc
T
Tpr = (1-12)
Tpc

————————
① psia refers to absolute pressure, 1psia = 6.985kPa.
·8· Petroleum Production Technology

where,
P = system pressure, psia;
T = system temperature, ºR;
Ppc = pseudo-critical pressure, psia;
Tpc = pseudo-critical temperature, ºR;
Ppr = pseudo-reduced pressure, psia;
Tpr = pseudo-reduced temperature, ºR.
(3) Find Z factor from Fig. 1-6. Several methods have been developed to directly
calculate gas Z factor. These methods are not introduced here.

Fig. 1-6 Gas Z factor chart


Chapter 1 Properties of Oil and Gas ·9·

1.2.2 Gas compressibility


The compressibility of gas is defined as the change in volume per unit volume for a
unit change in pressure. Gas compressibility is high, because gas is very compressible.
The gas compressibility is expressed as:
1 ⎛ ∂V ⎞
Cg = − ⎜ ⎟ (1-13)
V ⎝ ∂P ⎠T

Recall the gas equation of state,


nRTZ
V= (1-14)
P

Differentiating the gas equation of state with respect to pressure gives:


⎛ ∂V ⎞ ⎛ 1 ∂Z Z ⎞
⎜ ⎟ = nRT ⎜ − ⎟ (1-15)
⎝ ∂P ⎠T ⎝ P ∂P P 2 ⎠

Combining the derivative above and the definition of gas compressibility, we find:
1 1 ∂Z
Cg = − (1-16)
P Z ∂P

It is clear that we need to find the slope of Z at certain pressure for compressibility.
This approach is further illustrated in Fig. 1-7.

Fig. 1-7 Obtain gas compressibility from gas Z factor


·10· Petroleum Production Technology

1.2.3 Gas formation volume factor


The gas formation volume factor is defined as the ratio of gas volume at certain
pressure and temperature to that at standard conditions (60ºF and 14.7 psia).

VPT
Bg = (1-17)
Vsc

where,
Bg = gas formation volume factor;
VPT = volume of gas under certain pressure and temperature;
Vsc = volume of gas under standard conditions.
Applying the gas equation of state yields:
Psc ZT
Bg = (1-18)
Tsc P

For the standard conditions, the pressure is at 14.7 psia, and the temperature is at
520ºR. And the above Equation (1-19) becomes

ZT
Bg = 0.02827 (1-19)
P

1.2.4 Gas viscosity


Gas viscosity is affected by pressure, temperature, and gas composition. Under low
pressure, gas viscosity increases as temperature increases. Under high pressure, gas
viscosity decreases as temperature is raised. Gas viscosity can be measured in
laboratory. If measurement is not available, gas viscosity is often estimated with
available charts. Lee method can calculate gas viscosity, but this method can not be
used on sour gas unless the related factors have been corrected.

K ⎡ ⎛ ρ ⎞Y ⎤
μg = exp ⎢X ⎜ g ⎟ ⎥ (1-20)
104 ⎢ ⎝ 62.4 ⎠ ⎥
⎣ ⎦

where,
Chapter 1 Properties of Oil and Gas ·11·

(9.4 + 0.02M a )T 1.5


K=
209 + 19 M a + T

986
X = 3.5 + + 0.01M a
T

Y = 2.4 − 0.2 X

ρg = gas density under certain pressure and temperature, lb/ft3 ;


T = system temperature, ºR;
Ma = apparent molecular weight of gas mixture.
Example: A natural gas contains 85% Methane, 10% Ethane, and 5% Propane.
The gas is under 3000 psia and 200ºF. ① Find the apparent molecular weight and
specific gravity for the gas mixture. ② Find the Z factor for this gas mixture under the
system conditions ③ Calculate gas formation volume factor and gas viscosity under
system conditions.
Molecular weight of methane = 16
Molecular weight of ethane = 30
Molecular weight of propane = 44
Therefore,
Apparent molecular weight:
Ma = 0.85×16+0.1×30+0.05×44 = 13.6+3+2.2 = 18.8
Gas specific gravity = 18.8/29 = 0.648
Critical pressure for methane = 666.4 psia
Critical pressure for ethane = 706.5 psia
Critical pressure for propane = 616.4 psia
Pseudo-critical pressure:
Tpc = 0.85×666.4+0.1×706.5+0.05×616.4 = 667.9 (psia)
Critical temperature for methane = 343.3ºR
Critical temperature for ethane = 549.9ºR,
Critical temperature for propane = 666.1ºR
Pseudo-critical temperature:
Tpc = 0.85×343.3+0.1×549.9+0.05×666.1 = 380.1(ºR)
System temperature = 200+460 = 660(ºR)
Pseudo-reduced pressure = P/Ppc = 3000/667.9 = 4.5
Pseudo-reduced temperature = T/Tpc = 660/380.1 = 1.74
·12· Petroleum Production Technology

According to Fig. 9-4, we find Z factor is 0.92.


Gas formation volume factor:
ZT 0.92×659
Bg = 0.02827 = 0.02827 = 0.00571
P 3000

Gas density:
m PM 3000×18.8 lb
ρg = = = = 8.67 3
V ZRT 0.92×10.73×659 ft

Gas viscosity:

(9.4 + 0.02× M a )T 1.5 (9.4 + 0.02×18.8)×6591.5


K= = = 135
209 + 19× M a + T 209 + 19×18.8 + 659

986
X = 3.5 + + 0.01×18.8 = 5.183
T

Y = 2.4 − 0.2 X = 1.363

⎡ ⎛ ρ ⎞Y ⎤ 135 ⎡ 1.363 ⎤
K g ⎛ 8.67 ⎞
μg = 4 exp ⎢ X ⎜ ⎟ ⎥ = 4 exp ⎢5.183×⎜ ⎟ ⎥ = 0.019(cP)
10 ⎢ ⎝ 62.4 ⎠ ⎥ 10 ⎢
⎣ ⎝ 62.4 ⎠ ⎦⎥
⎣ ⎦

1.3 Properties of crude oil

Under reservoir pressure, significant amount of hydrocarbon is dissolved in crude oil.


When oil is produced, pressure and temperature reduce, and gas is liberated from the
oil phase. Changes in pressure, temperature and dissolved gas lead to changes in oil
properties. The most important oil properties include oil density, solution gas oil ratio,
oil formation volume factor, and oil viscosity.

1.3.1 Oil density


Oil specific gravity is defined as the ratio of crude oil density to that of water. Crude oil
specific gravity usually ranges from 0.8 to 0.9.
ρo
γo = (1-21)
ρw
Chapter 1 Properties of Oil and Gas ·13·

where,
γo = oil specific gravity;
ρo = density of oil;
ρw = density of water.
API gravity is often used in American oil industry. It is defined by the equation
below. API gravity often ranges from 50 for light oil to 10 for heavy oil and water.
141.5
API Gravity = − 131.5
γo

1.3.2 Solution gas oil ratio


Solution gas oil ratio (solution GOR, or gas solubility) is defined as the amount of gas
that evolves from oil as pressure is reduced. It often carries the unit of scf/stb (standard
cubic feet per stock tank barrel) or scm/t (standard cubic meter per metric ton). Its
behavior versus pressure is expressed in Fig. 1-8, where Pb is the bubble point pressure.

Fig. 1-8 Gas solubility versus pressure

When pressure reduces but still above bubble point pressure, no gas is liberated
and the initial solution GOR (Rsi) remains constant. After pressure drops below the
bubble point pressure, gas begins to liberate and solution GOR is reduced. In fact, any
pressure below original bubble point pressure is also a new bubble point pressure, since
oil is still saturated with gas at this point.
If the reservoir pressure remains above the bubble point, then any gas liberated
·14· Petroleum Production Technology

from the oil must be released in the tubing and the separators, and will therefore appear
at the surface. In this case, the producing GOR is equal to solution GOR, that is every
stock tank barrel of oil produced liberates Rs (standard cubic ft) of gas at surface.
If the reservoir pressure drops below the bubble point, then gas is liberated from
oil in the reservoir. The liberated gas may flow towards the producing wells due to the
lower pressure at the well, or it may migrate upwards to form a secondary gas cap
under the influence of the buoyancy force. Consequently, the producing GOR differs
from solution GOR.
In a saturated oil reservoir containing an initial gas cap, the producing GOR may
be significantly higher than the solution GOR of the oil, as free gas in the gas cap is
produced through the wells via a coning mechanism. Free gas is the gas existing in the
gas cap as a separate phase, distinct from solution gas which is dissolved in the oil
phase.
Gas solubility below bubble point pressure can be predicted by five various
correlations. Here only the Standing correlation is presented. Standing correlation is
based on California oil samples.

1.2048
⎡⎛ P ⎞ ⎤
Rs = γg ⎢⎜ + 1.4 ⎟10 x ⎥ (1-22)
⎣⎝ 18.2 ⎠ ⎦

x = 0.0125API − 0.00091T

where,
P = system pressure, psia;
T = system temperature, ºF;
γg = specific gravity for solution gas;
API = API gravity of oil.

1.3.3 Oil formation volume factor


The Oil formation volume factor (oil FVF or Bo)is defined as the ratio of the volume
of oil containing solution gas at the prevailing pressure and temperature, to the volume
of oil at standard conditions. When system pressure declines from initial reservoir
pressure, oil first expands and oil formation volume factor increases. After gas is
liberated below the bubble point pressure, oil shrinks and Bo decreases. This behavior
is plotted in Fig. 1-9. The value of oil formation volume factor depends on the fluid
Chapter 1 Properties of Oil and Gas ·15·

type and reservoir conditions. The value can vary from 1.1 for a black oil to 2.0 for
very volatile oil.

Fig. 1-9 Oil formation volume factor versus pressure

A few correlations are available for prediction of oil formation volume factor
under saturated conditions. Standing summarized a correlation for oil formation
volume factor based on 105 experiments for 22 California oil samples. This method
can only be used for oil at or below bubble point pressure.
1.2
⎡ ⎛ γ ⎞0.5 ⎤
g
Bo = 0.9759 + 0.00012 ⎢ Rs ⎜ ⎟ + 1.25T ⎥ (1-23)
⎢ ⎝ γo ⎠ ⎥
⎣ ⎦

where,
T = temperature, ºF;
γo = specific gravity of stock tank oil;
γg = specific gravity of solution gas.
Fig. 1-10 compares the solution GOR and oil formation volume factor predictions
with various methods. In practice, we need to go through a matching process to select
the best method.
·16· Petroleum Production Technology

Fig. 1-10 Predictions of parameters with various methods

1.3.4 Oil viscosity


Oil viscosity is strongly affected by oil composition, pressure, temperature, and gas
solubility. Oil viscosity is reduced at high temperature. Solution gas lightens oil and
leads to a reduction in oil viscosity.
Regarding oil viscosity, we should clarify dead oil viscosity, under-saturated oil
viscosity or saturated oil viscosity. Dead oil viscosity is the viscosity for oil at surface
conditions. The solution gas has escaped. Saturated oil viscosity is the viscosity of oil
at or below bubble point pressure. Under-saturated oil viscosity is measured at pressure
higher than bubble point pressure.
Oil viscosity should be measured in lab under certain pressure and temperature
when possible. Beggs and Robinson developed a correlation for dead oil viscosity
based on 460 measurements.

μod = 10 x − 1 (1-24)

where,
μod = dead oil viscosity, cP;
x = 10y T–1.163;
y = 3.0324–0.02023 API;
T = temperature, ºF.
Beggs and Robinson also developed a correlation for saturated oil viscosity based
Chapter 1 Properties of Oil and Gas ·17·

on 2073 measurements.
b
μos = aμod (1-25)

where,
μos = saturated oil viscosity, cP;
a = 10.715(Rs+100)–0.515;
b = 5.44(Rs+150)–0.338;
Rs = gas solubility, scf/stb.
Example: A reservoir oil is under 3000 psia and 200 ºF. The bubble point
pressure is 2500 psia. Oil specific gravity is 0.8, and gas specific gravity is 0.7. Predict
oil solution gas ratio and oil formation volume factor at 1000 psia and 100 ºF.
Solution:
Oil API gravity:
141.5 141.5
API Gravity = − 131.5 = − 131.5 = 45.38
γo 0.8

Oil solution gas ratio:


x = 0.0125API − 0.00091T = 0.0125×45.38 − 0.00091×100 = 0.476
1.2048 1.2048
⎡⎛ P ⎞ ⎤ ⎡⎛ 1000 ⎞ ⎤
Rs = γg ⎢⎜ + 1.4 ⎟10 x ⎥ = 0.7 ⎢⎜ + 1.4 ⎟×100.476 ⎥ = 337.3scf / stb
⎣⎝ 18.2 ⎠ ⎦ ⎣⎝ 18.2 ⎠ ⎦
Oil formation volume factor:
1.2
⎡ ⎛ γ ⎞0.5 ⎤
g
Bo = 0.9759 + 0.00012 ⋅ ⎢ Rs ⎜ ⎟ + 1.25T ⎥
⎢ ⎝ γo ⎠ ⎥
⎣ ⎦
1.2
⎡ ⎛ 0.7 ⎞
0.5 ⎤
= 0.9759 + 0.00012× ⎢337.3×⎜ ⎟ + 1.25×100 ⎥ = 1.155bbl / stb
⎢⎣ ⎝ 0.8 ⎠ ⎥⎦

1.4 Fluid sampling

Reservoir fluid sampling is usually done early in the field life in order to use the results
in the evaluation of the field and in the process facilities design. Once the field has
been produced and the reservoir pressure changes, the fluid properties will change as
described in the previous section. Early sampling is therefore an opportunity to collect
unaltered fluid samples.
·18· Petroleum Production Technology

Fluid samples may be collected down-hole at near-reservoir conditions, or at


surface. Subsurface samples are more expensive to collect, since they require
down-hole sampling tools. But subsurface sampling is more likely to capture a
representative sample, since a single-phase fluid is collected. A surface sample is
inevitably a two-phase sample which requires recombining to recreate the reservoir
fluid. When the pressure falls below the bubble point, both sampling techniques face
the same problem of trying to capture a representative sample (i.e. the correct
proportion of gas to oil).
Subsurface samples can be taken with a subsurface sampling chamber, called a
sampling bomb, or with a formation pressure testing tool, all of which are devices run
on wire line to the reservoir depth, as illustrated in Fig. 1-11.

Fig. 1-11 Fluid sampling operation

The sampling bomb requires the well to be flowing, and the flowing bottom hole
pressure is preferably above the bubble point pressure of the fluid to avoid phase
segregation. If this condition can be achieved, a sample of oil containing the correct
amount of gas is collected. If the reservoir pressure is close to the bubble point,
sampling should be done at low rates to maximize the sampling pressure. The valves
on the sampling bomb are open to allow the fluid to flow through the tool, and then
hydraulically or electrically close to trap a volume (typically 600 cm3) of fluid. This
small sample volume is one of the drawbacks of subsurface sampling.
Sampling saturated reservoirs with this technique requires special care to attempt
to obtain a representative sample, and in any case when the flowing bottom hole
pressure is lower than the bubble point, the validity of the sample remains doubtful.
Multiple subsurface samples are usually taken by running sample bombs in tandem or
Chapter 1 Properties of Oil and Gas ·19·

performing repeat runs. The samples are checked for consistency by measuring their
bubble point pressure at surface temperature. Samples whose bubble point lie within
2% of each other may be sent to the laboratory for PVT analysis.
Chapter 2 Properties of Reservoir Rock
For petroleum engineers, it is essential to understand rock properties that include
porosity, permeability, wettability, capillary pressure, relative permeability, and
electrical resistivity. But first, we need to understand the origin and the accumulation of
petroleum.

2.1 Basic petroleum geology

It is well documented that oil and gas accumulations are of organic origin and they
formed from organic matter in sediments. The organic matter from which petroleum is
derived originated through photosynthesis(i.e. storage of solar energy). Sunlight is
continuously transformed into such energy on earth but only a very small proportion of
the solar energy is preserved as organic matter and petroleum. Ninety percent of the
production of organic matter is from algae. Algae and to some extent zooplankton
account for the bulk of the organic material which can be transformed into oil. The
zooplankton and higher organisms that consumed the algae were thus indirectly
dependent on photosynthesis too. Larger animals such as dinosaurs are totally
irrelevant as sources of oil.
Basins with restricted water circulation will preserve more organic matter and
produce good source rocks which may mature to generate oil and gas. As organic
material becomes buried by the accumulation of overlying sediments, water is
gradually expelled during compaction.
Kerogen is a collective name for organic material that is insoluble in organic
solvents, water or oxidizing acids. The portion of the organic material soluble in
organic solvents is called bitumen, which is essentially oil in a solid state. Kerogen
consists of very large molecules. After long-term exposure to high temperature, the
large molecules crack into smaller molecules, mostly petroleum. In rapidly subsiding
basins, the exposure time is shorter and oil generation may only start at about 140℃ to
150℃. In the North Sea Basin, the oil window may typically be between 130℃ and 140℃.
Oil after deep burial over long time is cracked into gas.
Chapter 2 Properties of Reservoir Rock ·21·

Petroleum migrates from low permeability source rocks into high permeability
reservoir rocks. The main driving force for petroleum migration is buoyancy because
oil and gas are less dense than water. The factors acting against migration are the
capillary forces and low permeability. We distinguish between primary migration,
which is the flow of petroleum out of the source rock and secondary migration, which
is the continued flow from the source rock to the traps or up to the surface. The rate of
oil and gas formation in sedimentary basins each year is much less than the rate of
production and consumption. In practice petroleum is therefore regarded as a
non-renewable resource even though petroleum is being formed all the time.
Hydrocarbon traps are reservoir rocks overlain by tight (low permeability) cap
rocks that do not allow oil and gas to escape. The cap rock may not be 100% effective
in preventing the upward flow of hydrocarbons, but oil and gas still accumulate if the
rate of leakage is less than the rate of supply into the trap. Cap rocks are usually not
totally impermeable to water, but may be impermeable to oil and gas due to capillary
resistance in the small pores. Common traps include anticline trap, fault-controlled trap,
and salt domes, as depicted in Fig. 2-1.
·22· Petroleum Production Technology

Fig. 2-1 Anticline (a), fault (b) and salt dome (c) traps

Reservoir rocks are either of sandstone or carbonate composition. There are some
important differences between the two rock types which affect the quality of the
reservoir and its interaction with fluids which flow through them. The main component
of sandstone reservoirs is quartz (SiO2). Chemically it is a fairly stable mineral. It is
not easily altered by changes in pressure or temperature. Sandstone reservoirs form
after the sand grains have been transported over large distances and deposited.
Carbonate reservoir rock is usually found at the place of formation (in situ). Carbonate
rocks are susceptible to alteration by the processes of diagenesis.
The pores between the rock grains were initially filled with water. The migrating
hydrocarbons displace the water and thus gradually fill the reservoir. For a reservoir to
be effective, the porosity should be relatively high, and the pores not only need to have
good communication to allow fluid migration, but also allow flow towards the borehole
once a well is drilled into the structure.

2.2 Porosity

Rock porosity is defined as the ratio of pore volume to bulk volume. Most of the pores
are interconnected inside porous rocks, therefore fluid such as gas, oil and water can
flow through. But some pores are isolated and disconnected from other pores, due to
cementation and compaction. This phenomenon leads to two different types of porosity.
Absolute porosity takes into account all pore spaces, while effective porosity considers
only the connected pore spaces. It is obvious that effective porosity determines the
producible quantity of fluid in pore space.
Porosity is mainly determined by the packing of sand grains, as shown in Fig. 2-2.
Chapter 2 Properties of Reservoir Rock ·23·

For packed uniform spheres, the porosity of the structure is high. In reality, the porosity
of reservoir rock is lower, because the grains are of various shapes and sizes. On the
other hand, reservoir rock went through long-term cementation and compaction, which
also reduced porosity. Porosity of oil-bearing sandstones usually ranges from 5% to
25%. Porosity of limestone and dolomite is usually lower.

Fig. 2-2 Porosity of ideal structure and real rock

Rock porosity can be easily measured with saturation method. Dry rock sample is
saturated with a brine with known density. Its effective porosity can be calculated
based on the weight difference before and after saturation.
Example: A rock sample measures 3 cm in diameter and 5 cm in length. The dry
sample weighs 100 grams. The sample is then saturated with brine whose specific
gravity is 1.25. After saturation, the sample weighs 110 grams. Calculate the porosity
of the rock sample.
Diameter of sample = 3 cm
Length of sample = 5 cm
Bulk volume of sample = 0.25×π×32×5 = 35.34(cm3)
Mass of brine in the sample = 110 – 100 = 10(g)
Volume of brine in the sample = 10/1.25 = 8(cm3)
Porosity = 8/35.34 = 22.6%
Example: A rock sample measures 3 cm in diameter and 20% in porosity. Water
flows through the rock at a rate of 0.071 ml/min. Calculate the linear water velocity in
the rock.
Diameter of sample = 3 cm
Cross sectional area of the sample = 0.25×π×32 = 7.1(cm2)
·24· Petroleum Production Technology

The area open to flow is much smaller than the cross sectional area.
Fluid velocity = 0.071/(7.1×20%) = 0.05 cm/min
The pores in the rock are of various sizes, ranging from less than 1 micron to
several hundred microns. The fluid velocities inside different pore sizes are also
different. The example actually calculates the average fluid velocity in the rock.

2.3 Absolute permeability

Permeability indicates the rock’s capability to transmit fluids. It directly controls the
fluid flow rate in reservoir. Permeability was first defined by Henry Darcy in 1856. He
conducted experiments on flow of water through sand pack and summarized the
equation below, which is named Darcy’s law:
kA dP
Q= (2-1)
μ dL

or,
k dP
v= (2-2)
μ dL

where,
Q = flow rate through porous medium, cm3/s;
v = apparent fluid velocity, cm/s;
k = permeability of porous medium, D①;
A = cross sectional area of porous medium, cm2;
μ = viscosity of fluid, cP;
P = pressure, atm;
L = length of porous medium, cm.
The unit of permeability is Darcy. Most rocks have permeability less than 1 Darcy.
A unit milidarcy (mD) is thus defined, and 1 Darcy = 1000 mD. With field units,
Darcy’s law is expressed as:
kA dP
Q = 0.001127 (2-3)
μ dL

where,
Q = flow rate through porous medium, bbl/d;
————————
① 1D = 0.986923×10–12m2.
Chapter 2 Properties of Reservoir Rock ·25·

k = permeability of porous medium, mD;


A = cross sectional area of porous medium, ft2;
μ = viscosity of fluid, cP;
P = pressure, psi;
L = length of porous medium, ft.
If a porous rock is saturated with a single fluid, the permeability is only the
property of rock. The permeability at 100% saturation of a fluid is termed absolute
permeability. Absolute permeability is often conveniently measured by flowing brine or
dry gas, such as nitrogen through the porous medium. The test is illustrated in Fig. 2-3.
Nitrogen flows through the cylindrical core sample at a stable flow rate. The pressure
loss across the sample is measured. The rock permeability is calculated with Darcy’s
law.

Fig. 2-3 Measurement of rock permeability

Example: A rock sample measures 3 cm in diameter and 5 cm in length. A brine


is injected through the rock at a stable flow rate of 0.5 cm3/s. The brine viscosity is 1 cP
under testing temperature. The pressure difference across the sample is measured to be
2 atm. Calculate the permeability of the rock. Diameter of sample = 3 cm; Length of
sample = 5 cm.
Cross sectional area of the sample = 0.25×π×32 = 7.1(cm2)
According to Darcy’s law:
k = QμdL/(AdP) = 0.5×1×5/7.1×2 = 0.176 D = 176(mD)
Example: Find the nature of the unit Darcy with dimension analysis method.
Recall Darcy’s law:
k dP
v=
μ dL

Dimension of velocity = L/T (length/time)


·26· Petroleum Production Technology

Dimension of pressure = (M/L)/T2 (mass/length/time2)


Dimension of viscosity = (M/L)/T (mass/length/time)
Dimension of length = L (length)
We can find the dimension of the unit Darcy is actually area (L2).

Fig. 2-4 Linear flow through parallel permeable zones

Example: A well is drilled through three parallel permeable zones, as seen in Fig.
2-4. The thickness and permeability for the zones are given in the Tab. 2-1. Calculate
the average permeability. Assume the three zones have the same length and width.
Tab. 2-1 The thickness and permeability for zones
Zone Zone thickness/m Zone permeability/mD
1 10 200
2 20 100
3 15 300

For zone 1, applying Darcy’s law gives:


k1wh1 dP
Q1 =
μ dL

where,
Q1 = flow rate received by zone 1, cm3/s;
k1 = permeability of zone 1, D;
h1 = thickness of zone 1, cm;
w = width of rock, cm.
Similarly for the other two zones, we can write:
Chapter 2 Properties of Reservoir Rock ·27·

k2 wh2 dP
Q2 =
μ dL

and,
k3 wh3 dP
Q3 =
μ dL

For the three zones combined, applying Darcy’s law again yields:
ka w(h1 + h2 + h3 ) dP
Qt =
μ dL

where,
ka = average permeability, D;
Qt = total flow rate combined, cm3/s.
It is obvious that:

Qt = Q1 + Q2 + Q3

Therefore,
ka w(h1 + h2 + h3 ) dP k1wh1 dP k2 wh2 dP k3 wh3 dP
= + +
μ dL μ dL μ dL μ dL

Rearranging above equation yields:

ka (h1 + h2 + h3 ) = k1h1 + k2 h2 + k3 h3

The average permeability is expressed as:


k1h1 + k2 h2 + k3 h3
ka =
h1 + h2 + h3

or,

∑ ki hi
n

ka = 1 n
∑1hi
This average permeability is referred to as weighted-average permeability. Now
we calculate the average permeability with the data in the table.
·28· Petroleum Production Technology

200 × 10 + 100 × 20 + 300 × 15


ka = = 189mD
10 + 20 + 15

Fig. 2-5 Linear flow through sequential permeable zones

For the linear system presented in Fig. 2-5, the harmonic-average permeability can
be expressed as:

L1 + L2 + L3
ka =
L1 L2 L3
+ +
k1 k2 k3

or,

∑1 Li
n

ka =
nL
∑1 k i
i

where,
L1 = length of zone 1, ft or m;
k1 = permeability of zone 1, D.
Klinkenberg in 1941 discovered that permeability measurements made with gas as
the flowing fluid showed different results from permeability measurements made with
a liquid. Permeability measured with gas is always higher than that measured with
liquid. This phenomenon is named Klinkenberg effect.
It should be pointed out that rock properties vary vertically and laterally, which is
named reservoir heterogeneity. It has been proposed that most reservoirs are laid down
in a body of water by a long-term process. As a result of subsequent physical and
Chapter 2 Properties of Reservoir Rock ·29·

chemical reorganization, such as compaction, solution and cementation, the reservoir


characteristics are further changed. It is important to recognize homogeneous reservoirs
do not sexist in reality, only varying degrees of heterogeneity. In fact, reservoir
properties can vary dramatically at different locations. These properties include
porosity, permeability, reservoir thickness, faults and fractures, fluid saturations and
wettability.

2.4 Fluid Saturation


Saturation is defines as the fraction of pore volume occupied by a certain fluid (oil, gas,
or water). The oil saturation (So) can be expressed as:

volume of oil in pore space


So = (2-4)
pore volume

Similarly, gas saturation (Sg) and water saturation (Sw) can be defined. Even
though some reservoirs contain gas and some do not, it can be concluded that,

So + Sg + S w = 1 (2-5)

where,
So = oil saturation;
Sg = gas saturation;
Sw = water saturation.
Example: A rock sample measures 3 cm in diameter and 5 cm in length. Its
porosity is 20%. Initially the sample was saturated with water. Then 10 ml of oil was
slowly injected into the sample, and 4 ml of oil plus 6 ml of water was collected from
the outlet. Calculate the oil saturation in the rock sample.
Diameter of sample = 3 cm
Length of sample = 5 cm
Bulk volume of sample = 0.25×π×32×5 = 35.34(cm3)
Pore volume of sample = 35.34×20% = 7.1(cm3)
Volume of oil inside rock = 10–4 = 6(cm3)
Oil saturation = 6/7.1 = 84.5%
·30· Petroleum Production Technology

2.5 Surface tension and interfacial tension

Interfacial or surface tension exists when two or more phases are present. These phases
can be gas/oil, oil/water, or gas/water. For a liquid phase that is in contact with gas or
vapor, the force per unit length required to create a unit surface area is referred to as
surface tension. The surface tension of water ranges from 72.5 dyn/cm at 70ºF to 60
dyn/cm at 200ºF in an almost linear fashion. Dissolved natural gas reduces the surface
tension of water. At 70ºF, surface tensions of crude oils usually range from 24 to 38
dyn/cm. Dissolved gas greatly reduces the surface tension of oil. For a given oil,
surface tension decreases with increasing temperature.
Interfacial tension is the force that holds the surface of a particular phase together
and is normally measured in dyn/cm. The interfacial tension between water and crude
oil often ranges from 10 to 20 dynes/cm under reservoir conditions. It is a function of
pressure, temperature, and the composition of each phase. Interfacial tension reduces
with increasing temperature. General trend suggests lower interfacial tensions for low
density oils. Dissolved gas reduces interfacial tension.
At reservoir pressure and temperature, the interfacial tension between live oil and
water is higher than dead oil. Therefore, measurements should be made with live oil
rather than dead oil.
Surface tension and interfacial tension are important factors that govern the flow
of fluids in small pores in reservoir rocks. Fluid flow in a reservoir is affected by
surface tension, interfacial tension, pore sizes, pore shapes, and wettability of
fluid-rock system.

2.6 Wettability

The small pores in reservoir rocks have similar dimensions as capillary tubes. The flow
and distribution of fluids in rocks are affected by interfacial tension and rock
wettability. Wettability of a material is a measurement of the ability of a fluid to coat
the surface of the material. For example, water spreads on glass surface, so glass is
regarded as water-wet. Another example is that mercury forms droplets and fall off
glass surface, as seen in Fig. 2-6. In other words, mercury does not wet glass surface.
Chapter 2 Properties of Reservoir Rock ·31·

Fig. 2-6 Mercury on a glass surface

Wettability is quantified by contact angle. The contact angle (θ in Fig. 2-7) is the
angle at which the liquid-vapor interface meets the solid-liquid interface. As the
tendency of a drop to spread out over a flat solid surface increases, the contact angle
decreases. Thus, the contact angle provides an inverse measure of wettability.

Fig. 2-7 Contact Angle

A low contact angle that is less than 90° usually indicates that wetting of the
surface is very favorable, and the fluid will spread over a large area of the surface.
High contact angle (greater than 90°) generally indicates that wetting of the surface is
unfavorable, so the fluid will minimize contact with the surface and form a compact
liquid droplet. These tendencies are summarized in Tab. 2-2.
Tab. 2-2 Wettability based on contact angle
Contact Angle/(°) Wettability
θ=0 Spreading
θ < 90 Good wetting
θ = 90 or θ > 90 Incomplete wetting
θ > 180 Not wetting
·32· Petroleum Production Technology

In a reservoir, wettability is a measurement of the ability of a fluid to coat the rock


surface. In the earlier years, scientists believed pure silica should naturally be water
wet. However, subsequent research revealed wettability is a very complex issue. A
common method measures rock wettability with polished mineral crystals, which may
take several hundred hours to reach equilibrium. It was observed that the contact angle
changed with time. Some rock samples originally appeared to be water-wet, but turned
to be more oil-wet after long-term testing. It seems the polar contents in the oil
adsorbed to rock surface and altered wettability. Other wettability measurements
include Amott method and USBM method. The details of these methods are not
introduced here.
In fact, reservoir rocks can be water-wet, oil-wet, or intermediate-wet (also
known as neutral-wet). Rock wettability affects the oil and water distribution in the
pores, which is depicted in Fig. 2-8. In water-wet rock, a water layer wets the rock
surface and oil is located in the central parts of the pores. In oil-wet reservoir, oil coats
rock surface instead. If gas exists in the reservoir, gas is always the non-wetting phase.

Fig. 2-8 Rock wettability affects fluid distribution in porous media

On the other hand, the wettability at different locations in the same reservoir may
be different. Some portions of the reservoir are water-wet, but other portions are
oil-wet. This phenomenon is named fractional wettability. Another researcher proposed
mix-wettability, where the small pores are water-wet, but the large pores are more
Chapter 2 Properties of Reservoir Rock ·33·

likely to be oil-wet (Fig. 2-9). This explains some fields still produce oil at very low
oil saturation in the reservoir.

Fig. 2-9 Fluid distribution for different wettability

Based on contact angles, some criteria for wettability have been suggested by
scientists. These criteria are summarized in Tab. 2-2. For the 55 cores examined by
Treiber et al., 27% were found to be water-wet, 66% being oil-wet, and 7% being
intermediate-wet. If the Morrow criterion is applied, 26% were water-wet, 27% being
oil-wet, and 47% being intermediate-wet. Among the 161 carbonate cores tested by
Chilingar and Yen, 8% were water-wet, 65% being oil-wet, 15% being strongly oil-wet,
and 12% being intermediate-wet.
Tab. 2-3 Suggested contact-angle criteria for wettability
Contact angle Contact angle for Contact angle
Sources
for water-wet/(°) intermediate-wet/(°) for oil-wet/(°)

Treiber et al. 0~75 75~105 105~180

Morrow 0~62 62~133 133~180

Chilingar 100~160(oil-wet)
0~80 80~100
and Yen 160~180(strongly oil-wet)

2.7 Capillary action


Capillary action (sometimes named capillarity or capillary motion) is the ability of a
liquid to flow in narrow spaces without the assistance of, or even in opposition to,
external forces like gravity. The effect can be seen in the drawing up of liquids between
the hairs of a paint-brush, in a thin tube, in porous materials such as paper and plaster.
It occurs because of intermolecular forces between the wetting liquid and surrounding
·34· Petroleum Production Technology

solid surfaces. We study capillary action because the small rock pores have similar
dimensions as capillaries.
A common apparatus used to demonstrate the phenomenon is the capillary tube.
When the lower end of a vertical glass tube is placed in a liquid, such as water, a
concave meniscus forms, as shown in Fig. 2-10. Adhesion occurs between the fluid and
the solid inner wall pulling the liquid column up until there is a sufficient mass of
liquid for gravitational forces to overcome the intermolecular forces. On the other hand,
when glass tube is inserted into mercury, mercury level is pushed downwards and a
convex is formed.

Fig. 2-10 Capillary action of water and mercury

The height of a liquid column is given below. According to the equation, the
thinner the space in which the water can travel, the further up it goes.
2γ cos θ
h= (2-6)
ρgr

where,
γ = liquid-air surface tension;
θ = contact angle;
ρ = density of liquid;
g = local acceleration due to gravity;
r = radius of tube.

2.8 Capillary pressure

Capillary pressure is the pressure difference across the interface between two phases.
Chapter 2 Properties of Reservoir Rock ·35·

Similarly, it has been defined as the pressure differential between two immiscible fluid
phases occupying the same pores caused by interfacial tension between the two phases
that must be overcome to initiate flow.
Capillary pressure is important in reservoir engineering because it is a major
factor controlling the fluid displacement in a reservoir rock. Capillary pressure is only
observable in the presence of two immiscible fluids in contact with each other in
capillary tubes. The small pores in a reservoir rock are similar to capillary tubes, and
they usually contain two immiscible fluid phases in contact with each other. Thus
capillary pressure becomes an important aspect for reservoir rocks.
When two immiscible fluids are in contact with each other in a capillary tube, a clear
interface exists between them. This interface arises from interfacial tension effects. The
interface is a curved surface and the pressure on the concave side exceeds that in the
convex side. This pressure difference is known as capillary pressure, as illustrated in Fig.
2-11. In the presence of two immiscible fluids, one of them preferentially wets the tube
surface and it is called the wetting fluid, the other fluid is the non-wetting fluid. OWC
represents oil water contact.

Fig. 2-11 Capillary pressure

Capillary pressure can be mathematically expressed as:

Pc = Pnw – Pw (2-7)

where,
Pc = capillary pressure;
Pnw = pressure of the non-wetting fluid across the two-phase interface;
·36· Petroleum Production Technology

Pw = pressure of the wetting fluid across the two-phase interface.


An important expression relating the capillary pressure with the radius of the capillary
tube and the interfacial tension can be derived by balancing the pressures in the system
including the hydrostatic and interfacial tension. Now consider a system with oil and water,
where water is the wetting fluid:

2σ cos θ
Pc = (2-8)
r

and,

Pc = Po − Pw = ( ρw − ρo ) gh (2-9)

where,
σ = interfacial tension between two fluids;
Po = pressure of oil phase across interface;
Pw = pressure of water across interface.
In laboratory, capillary pressure is measured as a function of the saturation of the
wetting phase. In this sense, capillary pressure can be viewed as the necessary pressure to
force non-wetting fluid to displace the wetting fluid in a capillary. For example, a
water-wet reservoir rock is saturated with water and oil. Because of capillarity, the water
tends to rise inside the rock pores, due to the rock surface’s preference to this fluid. The
height at which the water rises in the rock depends on the capillary pressure between the
water and the oil. If the water is to be forced out, a pressure equal to the capillary pressure
must be applied in order to eject the water out of the capillary.
Rocks usually have pores of different radii. Therefore, a particular capillary pressure
will be associated with a specific set of pores having the same radius. Lower capillary
pressure will displace water out of bigger pores, but higher capillary pressures are required
to displace water out of smaller pores.
Now consider water as the wetting phase and oil as the non-wetting phase. Fig. 2-12
presents typical capillary pressure curves as a function of water saturation for a water-oil
system. Initially, the core is saturated with water. Oil (non-wetting phase) is injected into
the core to displace the water. As we can see from the primary drainage curve, a threshold
pressure (also know as entry pressure ) must be overcome before oil enters the core.
Drainage refers to the displacement of the wetting phase (water in this example) by the
non-wetting phase (oil in this case).
Chapter 2 Properties of Reservoir Rock ·37·

Fig. 2-12 Capillary pressure curves

As higher capillary pressure forces more oil to flow into the core, the water saturation
reduces until further increase in capillary pressure causes little or no change in water
saturation. This water saturation is named irreducible water saturation (Swi) or connate
water saturation (Swc).
It is common to plot two curves, one for the drainage process and one for the
imbibition process. Imbibition refers to the displacement of the non-wetting phase (oil in
this case) by the wetting phase (water in this case). For the imbibition process, there also
exists an oil saturation that can not be further reduced, namely residual oil saturation (Sor).
The imbibition curve presents lower capillary pressures for a fixed saturation than
drainage curves, which is named capillary hysteresis. Hysteresis is due to the following
factors: ① The advancing and receding contact angles between fluid and solid surface are
different. ② The wetting fluid has the natural tendency to saturate the rock. ③ Wettability
may change with time after contact with oil and water.
Capillary pressure curves can be generated with mercury injection method, restored
capillary pressure method, porous diaphragm method, or centrifuge method. Mercury
injection is the most widely-used method, while the restored capillary pressure method is
introduced here because it is the most accurate method. The apparatus is presented in Fig.
2-13. Firstly, saturate a core 100% with the reservoir water, then place the core on a
porous diaphragm or membrane which is saturated 100% with water and is permeable
to water only, under the pressure drops imposed during the experiment.
·38· Petroleum Production Technology

Fig. 2-13 Capillary pressure apparatus

Air is then admitted into the core chamber and the pressure is increased until a
small amount of water is displaced through the membrane into the graduated cylinder.
Pressure is held constant until no more water is displaced, which may require several
days or even several weeks. Afterwards, the core is removed from the apparatus, and
the water saturation is determined by weighing. The core is then again placed in the
apparatus, the pressure is increased, and the procedure is repeated until the water
saturation is reduced to a minimum.
Capillary pressure measurements serve very useful purposes. ① Pore radius
distribution can be obtained based on capillary pressure data, since contact angle and
interfacial tension remain constant. A steep slope on capillary curves indicates wide
pore size distribution. ② If capillary pressure data is corrected to reservoir conditions,
fluid saturation can be estimated from capillary pressures. The results often have good
agreement with logging data. ③ Capillary pressure data provides assessment for the
Chapter 2 Properties of Reservoir Rock ·39·

transition zone in the reservoir. A transition zone refers to the vertical thickness where
water saturation changes from 100% to connate water saturation at WOC (water-oil
contact), or from 100% liquid to Swc at GOC (gas-oil contact). The transition zone
must exist if the reservoir is supported by bottom water. WOC is defined as the
uppermost depth in the reservoir where 100% water saturation exists. GOC is defines
as the minimum depth where 100% liquid saturation exists.
Example: A well was drilled in Lulu field and capillary pressure was measured
with a core sample from the field. Besides, logging data reveals WOC is located at
5000 ft. Entry pressure on the primary drainage curve reads 2 psi, while the capillary
pressure at Swc reads 6 psi. Oil density and formation water density are 44 and 64 lb/ft3,
respectively. Interfacial tension between oil and water measures 50 dyn/cm. Based on
the available information, calculate: ① the depth of free water level; ② thickness of
transition zone.
The depth of free water level = 5000 + 144×2/(64–44) = 5014.4(ft)
Thickness of transition zone = 144×(6–2)/(64–44) = 28.8(ft)

2.9 Relative permeability

For multiphase flow in porous media, the relative permeability of a phase is a measure
of the effective permeability of a certain phase. It is the ratio of the effective
permeability of a certain phase to the absolute permeability. It can be viewed as an
adaptation of Darcy’s law to multiphase flow. Relative permeability is defined as:
ko
kro =
k

kg
krg =
k
kw
krw =
k

where,
ko = effective permeability of oil at a certain oil saturation;
kg = effective permeability of gas at a certain gas saturation;
kw = effective permeability of water at a certain water saturation;
kro = relative permeability of oil;
·40· Petroleum Production Technology

krg = relative permeability of gas;


krw = relative permeability of water.
Example: The absolute permeability of a rock sample is 100 mD. When the oil
saturation is 60%, the effective permeability to oil phase is 50 mD. Calculate the oil
relative permeability at this oil saturation.
ko 50mD
kro = = = 0.5
k 100mD

In a three-phase system, all three fluid phases coexist. As a fluid flows through the
porous media, it interferes with other fluids. Because capillary forces reduces the flow
rate of each individual phase in a non-linear fashion, consequently the sum of the
relative permeability of each phase is always less than one.
Typical oil-water two-phase relative permeability curves are presented in Fig.
2-14. As the water saturation increases, water relative permeability increases and oil
relative permeability decreases till residual oil saturation (Sor) is reached. The residual
oil saturation is the oil saturation below which the oil is immobile, that is, the oil
relative permeability is zero.

Fig. 2-14 Example of relative permeability curves

As the oil saturation increases, oil relative permeability increases and water
relative permeability decreases till connate water saturation (Swc) is reached. Connate
or irreducible water saturation is the water saturation below which water is not mobile
because of capillary forces. At connate water saturation, water remains static in the
smallest pores in the rock, as illustrated in Fig. 2-15. Therefore, the relative
permeability of water at connate water saturation is zero.
Chapter 2 Properties of Reservoir Rock ·41·

Fig. 2-15 Connate water reside in small pores

Wettability and heterogeneity have significant impacts on relative permeability.


For an oil-wet rock, its relative permeability curves show different features. The end
points are located differently. The cross point of two curves lean towards left side of
the chart. That is, the two curves cross at a water saturation less than 50%. An example
plot is given in Fig. 2-16.

Fig. 2-16 Relative permeability curves for oil-wet rock

An example of gas-oil relative permeability curves is presented in Fig. 2-17. Based


on the features of end points and cross point, it can bee seen gas is the non-wetting
phase. In fact, gas always serves as the non-wetting phase compared with water and oil.
·42· Petroleum Production Technology

Fig. 2-17 Relative permeability curves for gas and oil

In earlier years, scientists believed fluid viscosity had no effect on relative


permeability. However, later research showed fluid viscosity does affect relative
permeability when the displacing fluid is non-wetting. High viscosity of one fluid leads
to low relative permeability to the other fluid. The effect of temperature on relative
permeability is not yet clear.
A few correlations have been proposed to generate relative permeability curves based
on connate water saturation. The Corey method for gas-oil system is introduced here:
Sg
Sg* = (2-10)
1 − Swc

kro = (1 − Sg* ) 4 (2-11)

krg = ( Sg* )3 (2 − Sg* ) (2-12)

Example: Generate gas-oil relative permeability curves for connate water


saturation of 0.2. The calculation results are presented in Tab. 2-4. A plot can be made
based on these results.
Tab. 2-4 Relative permeability for gas and oil system
Sg Sg* kro krg
0.1 0.125 0.586 0.004
0.2 0.250 0.316 0.027
0.3 0.375 0.153 0.086
0.4 0.500 0.063 0.188
0.5 0.625 0.020 0.336
0.6 0.750 0.004 0.527
0.7 0.875 0 0.753
0.8 1 0 1
Chapter 2 Properties of Reservoir Rock ·43·

2.10 Electrical conductivity

Electrical conductivity is the ability of a material to transmit an electric current. Except


certain clay minerals, reservoir rocks do not conduct electricity. Oil, gas and pure water
do not conduct electricity either. Reservoir rocks containing water have good
conductivity, because large quantities of salts are dissolved in reservoir water.
Resistivity is the reciprocal of conductivity, as defined in the equation below.
A
R=r (2-13)
L

where,
R = Resistivity, Ω ⋅ m;
r = resistance of material, Ω;
A = cross sectional area of material, m2;
L = length of material, m.
The rock resistivity is of particular interests to logging engineers. Resistivity tools
measure formation resistivity by sending electric currents into the reservoir. When
resistivity log shows high resistivity reading, it may indicate existence of hydrocarbon.
An example is given in Fig. 2-18.

Fig. 2-18 High resistivity indicates hydrocarbon

Archie proposed the most important equation for resistivity logging:


·44· Petroleum Production Technology

aRw
Rt = (2-14)
φ m Swn
where,
a = tortuosity factor;
m = cementation factor/exponent;
n = saturation factor/exponent;
Rt = resistivity of rock containing fluids, Ω ⋅ m;
Rw = resistivity of formation water, Ω ⋅ m;
φ = porosity, %;
Sw = water saturation, %.
The tortuosity factor is meant to correct for variation in compaction, pore structure
and grain size. The parameter a is called the tortuosity factor and clearly is related to
the path length of the current flow. Its value lies in the range of 0.5 to 1.5.
The cementation factor models how much the pore network increases the resistivity, It
is related to the permeability of the rock. High rock permeability yields low cementation
exponent. For unconsolidated sands, the factor has been observed near 1.3. For
consolidated sandstones, common values are 1.8 to 2.0. In carbonate rocks, the cementation
exponent shows higher variance and values between 1.7 and 4.1 have been observed.
The saturation exponent usually is fixed to values close to 2. In logging
interpretation for a certain zone, rock porosity in Archie equation is obtained with
porosity logs. Resistivity of formation water is obtained by applying Archie equation to
a nearby zone that is saturated with formation brine only. Finally, the water saturation
and certainly the hydrocarbon saturation are obtained.
Various logging tools have been developed. Gamma ray log and spontaneous
potential log can locate the permeable zones. Induction log and lateral log indicate
existence of hydrocarbons. Neutron log, density log and sonic log measure the porosity
of formation. Readers can refer to a book on formation evaluation for details.

2.11 Rock sampling


Rock samples (namely cores) are required for direct measurements of physical rock
properties in a laboratory. They are also very useful for description of the depositional
environment, sedimentary features and the diagenetic history of the sequence.
Coring is performed in between drilling operations. Once the formation that
requires coring has been identified on the mud log, the drilling assembly is pulled out
Chapter 2 Properties of Reservoir Rock ·45·

of well. For coring operations, a special assembly is run on drill pipe comprising a core
bit and a core barrel, as shown in Fig. 2-19.

Fig. 2-19 Coring tool

Unlike a normal drill bit, a core bit is a hollow cylinder with an arrangement of
cutters on the outside. These cut a circular groove into the formation. Inside the groove
remains an intact cylinder of rock which moves into the inner core barrel as the coring
process progresses. Eventually, the core is cut free (broken) and prevented from falling
out of the barrel whilst being brought to surface by an arrangement of steel fingers or
catchers. Core diameters vary typically from 3 to 7 in(1in = 2.54 cm). and are usually
about 90 ft(1ft = 0.3048 m) long. However, in favorable conditions longer sections
may be achievable.
If a conventional core has been cut, it will be retrieved from the barrel on the rig
floor, cut into sections and crated. A lithologic description can be done at this stage. To
avoid drying out of core samples and the escape of light hydrocarbons some sections
will be immediately sealed in a coating of hot wax and foil.
In addition to a geological evaluation on a macroscopic and microscopic scale,
plugs (small cylinders of 3 cm diameter and 5 cm length) are cut from the whole core,
usually at about 30 cm intervals. Core analysis is carried out on these samples. Routine
·46· Petroleum Production Technology

core analysis includes measurements of porosity, horizontal permeability, fluid


saturation, and grain density. Special core analysis (SCAL) includes electrical tests,
relative permeability, capillary pressure, and strength tests.
Besides, sidewall sampling tool can be used to obtain small plugs (2 cm diameter,
5 cm length, often less) directly from the borehole wall. The tool is run on wire-line
after the hole has been drilled and logged. Some 20 to 30 individual bullets are fired
from each gun at different depths. The hollow bullet penetrates the formation and a
rock sample is trapped inside the steel cylinder (Fig. 2-20). By pulling the tool
upwards, wires connected to the gun pull the bullet and sample from the borehole wall.

Fig. 2-20 Sidewall sampling gun

Sidewall samples show hydrocarbons under UV light and differentiate between oil
and gas. The technique is applied extensively to sample microfossils and pollen for
stratigraphic analysis, such as age dating, correlation, depositional environment.
Qualitative inspection of porosity is possible, but very often the sampling process
results in a severe crushing of the sample, thus obscuring the true porosity and
permeability.
Chapter 2 Properties of Reservoir Rock ·47·

In a more recent development, a new wire-line tool has been developed that
actually drills a plug out of the borehole wall (Fig. 2-21). With sidewall coring, some
of the main disadvantages of the sidewall sampling gun are mitigated, in particular the
crushing of the sample. Up to 20 samples can be individually cut and stored in a
container inside the tool.

Fig. 2-21 Sidewall coring tool


Chapter 3 Reservoir Recovery Mechanisms
It is a very complex project to develop a commercial field, which involves economics,
technology and politics. In fact, about 90% of world reserves are owned and operated
by national oil companies, such as Saudi Aramco, Abu Dhabi National Oil Company
(ADNOC), Petroleos MexiCanos(PEMEX). Field exploration and production are
usually managed by one or more operators. Field services are provided by multiple
service companies.

3.1 Field life cycle

Firstly, an oil company evaluates the field of interest and gains access to the field. The
company needs to analyze the technical, economic, political, and environmental
challenges involved in field development. Technical analysis is focused on the potential
reserve, oil and gas properties (for example, the information on sulphur content in oil,
viscosity and density of crude oil, H2S content in natural gas, etc. if possible),
difficulties for exploration and production (for example, water depth if offshore).
Economic considerations include local taxation, local costs, repatriation of profits,
inflation, currency exchange rate, availability of skilled staff. Political considerations
include government political regime, civil and government stability. A company also
needs to consider the reputational damage if doing business in a country whose social
and political regime does not meet the approval of home government or share holders.
Moreover, a company must address the potential environmental risks. Independent oil
companies usually need to compete in public bidding and reach agreement with the
local government and local national oil company to develop an oil field.
The second phase is the exploration phase. For more than a century, geologists
have been looking for oil, and some major discoveries have been made. It is possible
that all the giant oil and gas fields have been explored. Future findings will be smaller
and more complex. Exploration techniques have improved significantly. However,
exploration remains a high-risk business. Exploration techniques include field work,
magnetic survey, gravity survey and seismic survey. It is common for a company to
carry out exploration work for some years before the first well is drilled.
Chapter 3 Reservoir Recovery Mechanisms ·49·

The third phase is the appraisal phase. Once the exploration well has encountered
hydrocarbon, necessary information and data are gathered for an initial estimation of
reserves. Afterwards, a feasible study is carried out to propose various plans for field
development. The study will contain the subsurface development options, the process
design, equipment sizes, the proposed locations for drilling and production sites, and
the crude evacuation and export system. The cases considered will be accompanied by
a cost estimate and planning schedule. Such a document gives a complete overview of
all the requirements, opportunities, risks and constraints.
The fourth phase is development planning. Based on the results of the feasibility
study, a field development plan can now be formulated and subsequently executed. The
plan is a key document used for achieving proper communication, discussion and
agreement on the activities required for the development of a new field, or extension to
an existing development. The plan should include the following information: Petroleum
engineering data, description of facilities, cost estimation, project planning, and budget
proposal. After the plan is approved, facilities and equipment will be designed,
fabricated, installed and commissioned.
The production phase commences when oil and gas reach surface. From now on,
cash flow is generated to pay back prior investments. The target of a new venture is to
minimize the time between exploration campaign and first oil. The field production is
greatly affected by the reservoir driving mechanisms. A field usually goes through
three phases. During the build-up phase, new well are drilled and production rate
increases. During the plateau phase, even though some new wells are drilled, but the
production of old wells declines. Thus the production remains relatively stable. Finally,
production from all wells begins to decline.
Once the net cash flow turns permanently negative, the economic life of the field
terminates and the field is decommissioned. It is of course still technically possible to
continue producing the field, but at a financial loss. Enhanced oil recovery methods
may be able to recover some proportion of the oil that still remains in the reservoir, if
the methods are cost-effective. Ultimately, all economically recoverable reserves will
be depleted and the field will be decommissioned. Decommissioning must minimize
the environmental effects without incurring excessive costs. Steel platforms may be cut
off to an agreed depth below sea level or toppled over in deep waters, whereas
concrete structures may be refloated, towed away and sunk in the deep ocean.
·50· Petroleum Production Technology

Pipelines may be flushed and left in place. In shallow tropical waters opportunities
may exist to use decommissioned platforms and jackets as artificial reefs in a
designated offshore area.

3.2 Primary recovery mechanisms

Fluids in reservoir require energy to flow through reservoir to reach surface. When the
reservoir is young, usually its pressure is high enough to lift fluids to surface, without
the aid of external energy. This process is referred to as primary recovery. Reservoir
behavior and performance are greatly affected by drive mechanisms. Primary drive
mechanisms include rock and liquid expansion drive, solution gas drive, gas cap drive,
aquifer drive, and combination drive.

3.2.1 Rock and liquid expansion drive


For a reservoir without initial gas cap or underlying active aquifer to support the
pressure, rock and liquid expand due to their compressibility while reservoir pressure
declines. This process leads to reduction in pore space. Because the compressibility of
rock and liquid is low, limited amount of fluids are driven out of reservoir. The
reservoir on expansion drive demonstrate rapid pressure decline and a constant GOR
(gas oil ration). Expansion drive is the least efficient drive mechanism that usually
yields less than 5% oil recovery.

3.2.2 Solution gas drive


Solution gas drive is also referred to as depletion drive, dissolved gas drive, and
internal gas drive. For a reservoir without initial gas cap or underlying active aquifer to
support the pressure, the gas originally dissolved in crude oil is liberated and expands
when reservoir pressure declines below bubble point pressure. The liberated gas forces
oil to flow. Under solution gas drive, reservoir pressure declines rapidly and
continuously. GOR climbs quickly while gas is liberated, but soon peaks due to
depletion of solution gas. In the absence of water aquifer, water production is limited.
The typical production history is depicted in Fig. 3-1.
Chapter 3 Reservoir Recovery Mechanisms ·51·

Fig. 3-1 Production profile for solution gas drive

The amount of solution gas is limited, thus solution gas drive mechanism is not
efficient. The typical recovery factor (RF in Fig. 3-1) under this drive is in the range
5%~30%, depending largely on the reservoir pressure, the solution GOR of the crude,
the abandonment conditions and the reservoir dip. Good recovery may be achieved by
a high dip reservoir that allows segregation of the secondary gas cap and the oil, a high
solution GOR, light crude and a high initial reservoir pressure.
In order to make use of the high compressibility of the gas, it is preferable that the
gas forms a secondary gas cap and contributes to the drive energy. This can be
encouraged by reducing the production rate per well. Producing wells should be
located away from the crest of the field. In a steeply dipping field, wells would be
located down-dip. However, in a field with low dip, the wells must be perforated as
low as possible to keep away from a secondary gas cap.
Field abandonment is usually due to high producing GOR and lack of reservoir
pressure to sustain production. For reservoirs under solution gas drive, secondary
recovery methods should be implemented at early stage.

3.2.3 Gas cap drive


For a gas cap reservoir, a gas zone lies above oil zone and no water aquifer is present.
As reservoir pressure declines, the gas cap expands and solution gas is liberated. These
two factors mobilize flow of oil to production wells. For a gas cap reservoir, production
·52· Petroleum Production Technology

wells should be completed away from oil-gas contact in order to delay gas coning, as
presented in Fig. 3-2.

Fig. 3-2 Well placement for gas cap reservoir

For reservoirs under gas cap drive, reservoir pressure declines at a relatively slow
rate, due to the energy provided by the highly compressible gas cap. Producing GOR is
stable before the gas cap reaches production wells’ completion zone. Afterwards, GOR
increases to very high values. Water-cut remains low because of negligible aquifer
movement. Typical production history is given in Fig. 3-3.

Fig. 3-3 Typical field production history under gas cap drive
Chapter 3 Reservoir Recovery Mechanisms ·53·

Gas cap drive yields relatively good recovery between 20% to 50%, influenced by
the field dip and the size of gas cap. Gas cap drive may be supplemented by reinjection
of produced gas, with the possible addition of make-up gas from an external source.
The gas injection well should be located in the crest of the structure, injecting into the
existing gas cap.

3.2.4 Aquifer drive


Many oil reservoirs are bounded by water aquifers. Some aquifers are so huge and
active that water effectively replaces the space from extraction of oil. The aquifer drive
(also known as natural water drive) mechanism is depicted in Fig. 3-4. Since the water
compressibility is low, the volume of water must be large to make this process effective,
hence the need for the large connected aquifer. In this context, large aquifer means 10
to 100 times the volume of oil in place.

Fig. 3-4 Aquifer drive mechanism

With active aquifer support, reservoir pressure declines very slowly and GOR is
stable, as show in Fig. 3-5. However, water may break through at early stage and water
production increases quickly. In this case, some wells may require water-shutoff
procedures or abandonment due to excessive water production. Active aquifer drive
yields very good recovery from 30% up to 70%.
·54· Petroleum Production Technology

Fig. 3-5 Field production history under aquifer drive

The prediction of the size and permeability of the aquifer is usually difficult, since
exploration and appraisal wells are usually targeted at locating oil, and typically little
data is collected in the water column. Hence the prediction of aquifer response often
remains a major uncertainty during reservoir development planning.
Water drive may be imposed by water injection into the reservoir, preferably by
injecting into the water column. If the permeability in the water leg is too low, it may
be necessary to inject into the oil column. Once water injection is adopted, the effect of
natural aquifer is usually negated. Clearly if it were possible to predict the natural
aquifer response at the development planning stage, the decision to install water
injection facilities would be made easier. A common solution is to initially produce the
reservoir using natural depletion, and to install water injection facilities in the event of
little aquifer support.

3.2.5 Combination drive


Some reservoirs are bounded with a gas cap and aquifer, as shown in Fig. 3-6. The
fluid flow is driven by all the mechanisms mentioned earlier. Due to the many factors
involved, the production profile often displays mixed signals. Many wells experience
early gas breakthrough, as well as increased water production. Excessive gas and water
production impact oil recovery negatively.
Chapter 3 Reservoir Recovery Mechanisms ·55·

Fig. 3-6 Reservoir under combination drive mechanism

A survey was done on the oil recovery for various fields under solution gas drive
and aquifer drive. The results are summarized in Tab. 3-1. It is obvious aquifer drive
yields much better oil recovery. In Tab. 3-1, OOIP refers to original oil in place.
Tab. 3-1 Primary recovery under two drive mechanisms
Primary recovery/%
Drive mechanism Lithology Location
(OOIP)

Solution gas drive Sandstone California 22


Solution gas drive Sandstone Louisiana 27
Solution gas drive Sandstone Texas 7C,8,10 15
Solution gas drive Sandstone Texas 1-7B,9 31
Solution gas drive Sandstone West Virginia 21
Solution gas drive Sandstone Wyoming 25
Solution gas drive Carbonate All 18
Aquifer drive Sandstone California 36
Aquifer drive Sandstone Louisiana 60
Aquifer drive Sandstone Texas 54
Aquifer drive Sandstone Wyoming 36
Aquifer drive Carbonate All 44

3.3 Secondary recovery methods

After some years of production, reservoir pressure has declined and production rate has
reduced. To improve production, water and gas are often injected into reservoirs to
maintain reservoir pressure and drive reservoir fluids. These operations, namely gas
flooding and water flooding, are regarded as secondary recovery methods. Water
flooding is the most widely-used secondary recovery method, adopted by many
onshore and offshore fields in the world.
·56· Petroleum Production Technology

The principal purpose for water flooding an oil reservoir is to increase the oil
production rate and ultimately, the oil recovery. This is accomplished by injection of
water to increase the reservoir pressure to its initial level and maintain it near that
pressure. Moreover, water drives oil to production wells like a piston, as shown in Fig.
3-7. Water flooding is the most popular secondary recovery method because water is
inexpensive and generally readily available in large quantities from nearby streams,
rivers, or oceans. Water injection effectively helped production wells near the water
injection wells to produce at higher rates because of the increased reservoir pressure.

Fig. 3-7 Water flooding process

Saline water or formation brine is produced from a well along with oil. As the oil
production rate declined, the water production rate often increases. In the early days of
the oil industry, the produced water typically was disposed of by dumping it into
nearby streams or rivers. In the 1920s, the practice began of re-injecting the produced
water into subsurface formations, including the oil reservoir from which the oil and
water originally had come. By the 1930s, produced water re-injection (PWRI) had
become a common oilfield practice.
Reinjection of water was first done systematically in the Bradford oil field of
Pennsylvania. There, a line flood pattern was used, in which two rows of producing
wells were staggered on both sides of an equally spaced row of water-injection wells.
In the 1920s, besides the line flood, a five-spot well layout was used.
Common injection patterns are direct line drive, staggered line drive, two-spot,
three-spot, four-spot, five-spot, seven-spot and nine-spot, as seen in Fig. 3-8. Normally, the
two-spot and three-spot patterns are used for pilot testing purposes. The patterns are called
Chapter 3 Reservoir Recovery Mechanisms ·57·

normal or regular when they include only one production well per pattern. Patterns are
described as inverted when they include only one injection well per pattern.

Fig. 3-8 Well patterns for water flooding

By combining Darcy’s law with expression of capillary pressure, Leverett


proposed the equation for fractional flow of water in the reservoir. For horizontal
displacement of oil by water, the simplified equation states:
1
fw = (3-1)
kro μw
1+
krw μo

The fraction of water (or water-cut) in the flowing stream is defined as:
qw
fw = (3-2)
qw + qo
·58· Petroleum Production Technology

where,
fw = fraction of water or water-cut;
qw = water flow rate in situ;
qo = oil flow rate in situ.
According to Leverett equation for fractional flow, a relation between water
fraction and water saturation can be plotted with relative permeability data. An
example is presented in Fig. 3-9.

Fig. 3-9 Relations of water fraction and water saturation

Buckley and Leverett later presented an equation to describe movement of water


front in the reservoir:
q ∂f w
v= (3-3)
Aφ ∂S w

where,
v = velocity of water front, m/d;
q = flow rate, m3/d;
A = reservoir cross sectional area, m2;
φ = rock porosity.
This equation states that the velocity of a front of constant water saturation is
Chapter 3 Reservoir Recovery Mechanisms ·59·

directly proportional to the derivative of water-cut at that saturation. Welge presented a


simple graphic solution, as shown in Fig. 3-10. First, plot the relation of fw versus Sw.
Second, draw a tangent line from initial water saturation (Swi). The tangent point
reveals the water saturation at the front (Swf). Third, extrapolate the tangent line till fw
equals 1 to find the average water saturation in the water bank.

Fig. 3-10 Welge plot for water front saturation

The oil recovery achieved by water flooding is greatly influenced by fluids


mobility ratio. In a porous medium, the mobility of a fluid is defined as its endpoint
permeability divided by its viscosity. That is, a fluid with a low viscosity has a high
mobility unless its effective permeability is very low. Similarly, a high-viscosity oil has
low mobility. For example, mobility of oil is expressed as:
k
λ = kro (3-4)
μo

where,
λ = mobility, mD/cP①;
kro = relative permeability of oil;
————————
① 1cP = 10–3Pa ⋅ s.
·60· Petroleum Production Technology

k = rock absolute permeability, mD;


μo = oil viscosity, cP.
When water displaces oil in the reservoir, the water-oil mobility ratio determines
which phase flows faster in the pore space. The mobility ratio is defined as:
krw μw krw μo
M= = (3-5)
kro μo kro μw

where,
krw = relative permeability of water;
μw = water viscosity, cP.
Because water’s viscosity at reservoir temperatures is usually much lower than
that of the reservoir oil, the water-oil mobility ratio is often much greater than 1.
Mobility ration greater than 1 indicates unfavorable displacement and water travels
faster than oil without effectively sweeping oil. This phenomenon is referred to as
viscous fingering, as illustrated in Fig. 3-11. As a result, water breaks through early at
production wells.

Fig. 3-11 Mobility ratio affects fluid displacement

The water-oil mobility ratio is a key parameter in determining the efficiency of the
water-oil displacement process. The recovery factor increases as the water-oil mobility
ratio decreases, as shown in Fig. 3-12.

Fig. 3-12 Mobility ratio versus recovery factor


Chapter 3 Reservoir Recovery Mechanisms ·61·

For some low permeability reservoirs, the injection of water is difficult and gas
flooding may be feasible. Gas flooding takes place as either a miscible flood or an
immiscible flood. Immiscible flooding means that the gas phase injected into the
reservoir does not mix with the oil phase in reservoir. Therefore the purpose of the
immiscible flood is to provide the energy or drive by increased pressure. Immiscible
gas flooding is considered a secondary recovery method.
In previous gas flooding projects, natural gas and nitrogen have been injected.
Because the gas mobility is very high, it easily channels through reservoir without
effectively sweeping oil. An alternative approach is foam flooding. Foam is produced
by mixing water, gas and detergent. Foam is more rigid than gas only. Water alternating
gas (WAG) has also been proved effective for a better sweep efficiency. For a WAG
process, gas is injected then followed by injection of water slug which reduces the
mobility of injected gas.
The recovery factors for some fields are presented in Tab. 3-2. Most fields
received water injection. Two fields that received edge water injection yield high
recovery factors around 55%. Gas injection into the gas cap yields oil recovery about
44%. Nevertheless, large quantities of crude oil is left behind after primary and
secondary recovery phases. Various methods have been developed to squeeze some
extra oil out of the reservoir.
Tab. 3-2 Recovery factors for various fields
Primary oil Secondary oil
Flooding mechanism Lithology
recovery/%(OOIP) recovery/%(OOIP)

Pattern water flood California sandstone 26.5 8.8

Pattern water flood Louisiana sandstone 36.5 14.7

Pattern water flood Oklahoma sandstone 17.0 10.6

Pattern water flood Texas sandstone 25.6 12.8

Pattern water flood Wyoming sandstone 23.6 21.1

Pattern water flood Texas carbonate 15.5 16.3

Edge water injection Louisiana sandstone 41.3 13.8

Edge water injection Texas sandstone 34.0 21.6

Gas injection into cap California sandstone 29.4 14.2

Gas injection into cap Texas sandstone 35.3 8.0


·62· Petroleum Production Technology

3.4 Enhanced oil recovery

Over the years, a few methods have been developed to further improve oil recovery by
injection of certain foreign agents into reservoirs. Common agents for injection include,
but are not limited to, steam, polymer, surfactants, alkaline solution, and microbes.
These methods are referred to as tertiary recovery methods or enhanced oil recovery
methods. They can be grouped into four categories as follows.

3.4.1 Thermal methods


Thermal methods include steam flooding and internal combustion. They are most
effective for heavy oil reservoirs. The internal combustion process is not easy to
control, therefore its field applications are limited. Among the many EOR methods,
steam flooding is the only routinely used method on a commercial basis. Many
relatively shallow heavy-oil reservoirs with high oil saturation are good candidates for
steam flooding. Otherwise, it is very difficult to mobilize the heavy oil without extra
heat. There are two common methods for steam injection, namely cyclic steam
injection and steam flooding.
Cyclic steam injection, also known as steam huff and puff, consists of 3 stages:
injection, soaking, and production. The mechanism proceeds through cycles of steam
injection, soak, and oil production. First, steam is injected into a well at a temperature
of 300℃ to 340℃ for a period of weeks to months. Next, the well is allowed to sit for
days to weeks to allow heat to soak into the formation. Finally, the hot oil is pumped
out of the well for a period of weeks or months.
Once the production rate falls off, the well is put through another cycle of
injection, soak and production. This process is repeated until the cost of injecting steam
becomes higher than the profits made from producing oil. This method can yield
recovery factors around 20% to 25% but the cost to inject steam is high. It is quite
common for wells to be produced in the cyclic steam manner for a few cycles before
being put on a steam flooding regime with other wells.
In a steam flood, sometimes known as a steam drive, some wells are used as steam
injection wells and other wells are used for oil production. Two mechanisms are at
work to improve the amount of oil recovered. The first is to heat the oil to higher
temperatures and to thereby decrease its viscosity so that it flows more easily through
Chapter 3 Reservoir Recovery Mechanisms ·63·

the formation toward the producing wells. A second mechanism is the physical
displacement in a manner similar to water flooding. While more steam is needed for this
method than for the cyclic method, it is typically more effective at recovering a larger
portion of the oil. Because of reservoir heterogeneity, a highly permeable zone with
low oil saturation often exists between the production well and injection well. In this
case, the injected steam channels through the high-permeability zone, without
effectively heating up the oil in place. While designing a steam flood, the criteria in
Tab. 3-3 can be used as guidelines. A form of steam flooding that has become popular
in the Alberta oil sands is steam assisted gravity drainage (SAGD), in which two
horizontal wells are drilled, one a few meters above the other, and steam is injected into
the upper one. The intent is to reduce the viscosity of the bitumen to the point where
gravity will pull it down into the producing well.
Tab. 3-3 Screening criteria for steam flood
Reservoir
Design Oil API In-situ oil Pay zone Reservoir Oil Reservoir Rock
Depth average
factors gravity viscosity thickness porosity saturation pressure lithology
permeability
Less
Application Less than More More than More More Less than
10 to 25 than Sandstone
criteria 15000 cP than 20 ft 200 mD than 20% than 40% 1500 psi
5000 ft

3.4.2 Miscible gas flood


Previously, immiscible gas flooding was introduced as a secondary recovery
method. Miscible gas flooding is considered an EOR method. Miscible gas flooding
typically employs CO2, natural gas, N2, or flue gas (87% N2, plus 12% CO2) for
injection. Miscible means that the injected gas mixes with the oil phase in place,
thereby reducing oil viscosity and mobilizing oil flow. On the other hand, the oil phase
swells after mixing with the injected gas.
When designing a miscible gas flood, it is crucial to obtain the minimum
miscibility pressure (MMP). Minimum miscibility pressure is generally used as the
key criteria for evaluating and selecting the most suitable solvents for improved oil
recovery projects. This is the lowest pressure by which gas can attain miscibility via a
multi-contact process at reservoir temperature in the given oil reservoir. The oil
reservoir for miscible gas flooding should be operated above or at the minimum
miscibility pressure. The parameter, before field trial, must be measured at laboratory
with rising bubble or slim tube experiment.
·64· Petroleum Production Technology

Fig. 3-13 presents the effect of pressure and oil composition on MMP. A high
pressure is required to achieve miscibility for a reservoir under high temperature. On
the other hand, a higher pressure is required when the oil phase contains more heavy
components.

Fig. 3-13 Factors that affect MMP

N2 is cheaper than CO2, but requires a much higher pressure to reach miscibility.
The MMP for N2 is 4~5 times higher than that for CO2, which limits its field
applications. As such, even though various gases have been injected for a better
recovery, the most favorable gas for injection is CO2, as long as the supply of CO2 is
available.
Today, most of the CO2 used in EOR operations is from natural underground
domes of CO2. With the natural supply of CO2 limited, man-made CO2 from the
captured CO2 emissions of power plants and industrial facilities (e.g., fertilizer
production, ethanol production, cement and steel plants) can be used to boost oil
production through EOR. Once CO2 is captured from these facilities, it is compressed
and transported by pipeline to oil fields. Several large scale CO2 storage and EOR
projects are being carried out in North America and the results are promising.
Chapter 3 Reservoir Recovery Mechanisms ·65·

Sask Power's boundary dam project retrofitted its coal-fired power station in 2014
with carbon capture and sequestration technology. The plant will capture 1 million
tonnes of CO2 annually, which is sold to Cenovus Energy for enhanced oil recovery at
the Weyburn oil field. The project is expected to recover an additional 130 million
barrels (21 million m3) of oil, extending the life of the oil field by 25 years. It is
projected that more than 26 million tonnes of CO2 will be stored in Weyburn, plus
another 8.5 million tonnes stored at the Weyburn-Midale Carbon Dioxide Project.
The resulting net reduction in atmospheric CO2 by CO2 storage in the oil field is
equivalent of taking nearly 7 million cars off the road for a year. Since CO2 injection
began in late 2000, the EOR project has performed largely as predicted. Currently,
some 1600 m3 (or 10063 barrels) per day of incremental oil is being produced from
the field.
Mississippi Power's Kemper county energy facility, or Kemper project, is a
first-of-its-kind plant in the U.S. and came online in 2016. The Southern Company
subsidiary has worked with the U.S. Department of Energy and other partners to
develop cleaner, less expensive, more reliable methods for producing electricity with
lignite and coal that also support oil production with CO2 injection.
For the Kemper project, rather than burning coal directly to make electricity,
gasification technology breaks down the coal into chemical components, removes
impurities before it is fired, avoids certain emissions, and takes gases that result from
this chemical breakdown to fuel the power plant. This plant is more efficient and
therefore cleaner than traditional coal-fired power plants. Additionally, the unique
location of the Kemper project, and its proximity to oil fields, makes it an ideal
candidate for enhanced oil recovery projects.
The main barrier to taking further advantage of CO2 EOR in the United States has
been an insufficient supply of affordable CO2. Currently, there is a cost gap between
what an oilfield operator could afford to pay for CO2 under normal market conditions
and the cost to capture and transport CO2 from power plants and industrial sources. As
a result, most CO2 comes from natural sources. However, using CO2 from power plants
or industrial sources could reduce the carbon footprint when the CO2 is stored
underground.
For some industrial sources, such as natural gas processing or fertilizer and
ethanol production, the cost gap is small (potentially $10~20/t CO2). For other
sources of CO2, including power generation and a variety of industrial processes,
·66· Petroleum Production Technology

capture costs are greater, and the cost gap becomes much larger (potentially $30~50/t
CO2). The enhanced oil recovery initiative has brought together leaders from industry,
the environmental community, labor, and state governments to advance CO2 EOR in
the United States and close the price gap. While designing CO2 flooding, the following
criteria in Tab. 3-4 can be useful.
Tab. 3-4 Screening criteria for miscible CO2 flood
Design parameters Application criteria
Oil API gravity More than 26
In-situ oil viscosity Less than 15 cP
Reservoir depth More than 2000 ft
Pay zone thickness Wide range
Oil saturation More than 30%
Porosity Wide range
Reservoir pressure, psi More than MMP
Rock type Sandstone or carbonate

3.4.3 Chemical methods


Chemical flooding methods involve injection of polymers, surfactant and alkaline
solutions into reservoirs, chased by routine water injection. In practice, these chemicals
are mixed with water prior to injection.
In an ideal water flooding process, water displaces oil like a piston. However, due
to reservoir heterogeneity and high water mobility, water fingers and channels through
reservoir without effectively displacing oil, as shown in Fig. 3-14. As a result, the
recovery factor is low. Polymer flooding is usually applied when the mobility ratio is
high or the heterogeneity of the reservoir is high. Polymer flooding method has
achieved large-scale field applications in China since 1990s.

Fig. 3-14 Mechanism of polymer flood


Chapter 3 Reservoir Recovery Mechanisms ·67·

According to Fig. 3-12, recovery factor improves when mobility ratio reduces.
When conducting a polymer water flood, a high molecular weight polymer is added to
the water of the water flood, to increase water viscosity and to decrease the mobility of
the flood water. As a consequence, the sweep efficiency of the water flood is improved.
Moreover, polymer flood yields better recovery by contacting more reservoir volume.
Thus, polymer flood can lead to ultimate recovery factor that is 4%~10% higher than
with plain water flooding at economic limit. Additionally, less water injection is required
to produce a given amount of oil. It is generally believed that polymer flood can not reduce
residual oil saturation. However, some research from China disputes this conclusion.
In Chinese oil fields, the most widely used polymer for commercial polymer
flooding projects is partially hydrolyzed polyacrylamide (HPAM) with molecular
weight above 20 million. Its structure is given in Fig. 3-15. HPAM has good ability to
increase water viscosity, but the solution degrades under high temperature and high
salinity, as seen in Fig. 3-16. HPAM is not recommended for reservoir of temperature
above 85ºC and high salinity.

Fig. 3-15 Molecular structure of HPAM

Fig. 3-16 Polymer viscosity under shearing and high salinity


·68· Petroleum Production Technology

For example, polymer flood with HPAM achieved good results in Daqing field in
China, thanks to the low reservoir temperature (around 45℃), and low salinity of
produced water (only 4000 mg/L). Certain biopolymers, such as xanthan gum, welan
gum and schizophyllan are more stable under high temperature and high salinity. But
the high prices limit their field applications.
The primary purpose of adding polymer to water is to increase the viscosity of the
flood water; however, it was observed that HPAM in many instances altered flow paths
by reducing the permeability to water. The reduction of permeability to water induced
by HPAM can last long time, while the permeability to oil remains relatively unchanged.
The effect of permeability reduction is quantified with resistance factor and residual
resistance factor. The resistance factor is defined as the ratio of the mobility of brine to
that of polymer fluid. It is thus expressed as:

λw krw /μw krw μp


RF = = = (3-6)
λp krp / μp krp μw

where,
RF = resistance factor;
krp = relative permeability of polymer fluid;
μp = polymer viscosity, cP.
The residual resistance factor is defined as the ratio of rock permeability to brine
before polymer flooding and that after polymer flooding. It is expressed as:
kb
RRF = (3-7)
ka

where,
RRF = residual resistance factor;
kb = permeability to brine before polymer injection, mD;
ka = permeability to brine after polymer injection, mD.
Polymer retention will often profoundly affect the technical and economic success
of a polymer flooding project. Excessive polymer retention leads to excessive costs of
polymer and reduced profits. Polymer retention includes adsorption, mechanical
trapping, and hydrodynamic retention. Mechanical entrapment and hydrodynamic
retention occur only to flowing polymer solution in porous media. Retention by
mechanical entrapment occurs when larger polymer molecules become lodged in
Chapter 3 Reservoir Recovery Mechanisms ·69·

narrow pore channels. Adsorption refers to the interaction between polymer molecules
and the solid surface. This interaction causes polymer molecules to be bound to the
surface of the solid, mainly by physical adsorption, van der Waals forces, and hydrogen
bonding. Mechanical trapping and adsorption contribute to majority of polymer
retention in porous media.
A proper polymer flood usually requires injection of 15% to 25% of reservoir pore
volume. Polymer concentrations often range from 250 to 2000 mg/l. For very large
field project, over one million kilograms of polymers can be consumed over a 1~2 years
period of time. While designing polymer flood, the criteria in Tab. 3-5 should be
considered.
Tab. 3-5 Screening criteria for polymer flood
Design parameters Application criteria
Oil API gravity More than 15
In-situ oil viscosity 10 to 150 cP
Oil saturation Mobile oil more than 10% pore volume
Depth Less than 9000 ft
Pay zone thickness Wide range
Permeability More than 800 mD preferred
Reservoir temperature Lower than 70℃ preferred
Rock type Sandstone preferred

Surfactants and alkali can be added to polymer solution for a better oil recovery.
Surfactants are chemical substances that concentrate on liquid/liquid interfaces at low
concentration, greatly reducing interfacial tension between fluids. It has been proved
that a relation exits between oil recovery and capillary number that is defined as
follows.

N ca =
σ

where,
Nca = capillary number;
v = fluid velocity, m/s;
μ = fluid viscosity, kg/ms;
σ = interfacial tension, N/m.
·70· Petroleum Production Technology

In essence, the capillary number is the ratio of viscous versus capillary forces. A
small capillary number suggests that the motion of the fluid is dominated by capillary
forces. Conversely, a large capillary number indicates a viscous dominated regime.
From a practical point of view, we wish to increase the capillary number thereby
reducing trapping. In practice, capillary number needs an increase by three orders of
magnitude to halve the residual oil saturation. This means a better recovery is achieved
at ultralow interfacial tension thanks to addition of surfactants (Fig. 3-17).

Fig. 3-17 Capillary number versus residual oil saturation

While alkali is added to the flood water, alkali reacts with the organic acids in
crude oil and produces surfactants, thus resulting in low interfacial tension and better
oil recovery. But alkali must be compatible with formation water. For instance,
alkali-surfactant-polymer (ASP) flood was abandoned at Shengli field due to severe
scale problem.
Chapter 4 Fluid Flow in Reservoir
Fluid flow in porous media is a very complex phenomenon. Fluid flows through the
tiny pores of non-uniform shapes and sizes, even though the pores are often simulated
as capillaries numerically. Moreover, two phases or even three phases flow
simultaneously in reservoir rock. As such, fluid properties, phase saturations, and
relative permeability play important roles in reservoir fluid flow. On top of these
factors, flow regime and reservoir geometry contribute extra difficulties to this
complex issue.

4.1 Basic concepts


In general, reservoir fluids are classified into three types: ① incompressible fluid. The
volume (or density) of incompressible fluids does not change with pressure. In reality,
incompressible fluids do not exist, but in some cases fluids can be assumed as
incompressible to simplify certain flow equations. ② slightly compressible fluids.
These fluids, such as crude oil and water, exhibit small changes in volume and density
with changes in pressure. ③ compressible fluids experience significant changes in
volume with changes in pressure. All gases are regarded as compressible fluids.
After the field is put on production, the reservoir pressure can demonstrate
different flow regimes: ① steady state. Under this scenario, the pressure at every
location in reservoir remains constant. Most reservoirs experience some level of
pressure decline while fluids are being extracted. However, if the reservoir is supported
by a large and active aquifer, or adequate water is injected into the reservoir for
purpose of better recovery, the void space from oil extraction is recharged by water.
Therefore, the reservoir pressure remains relatively constant due to aquifer support or
water injection. ② pseudo-steady state refers to the fact that reservoir pressure
declines linearly as a function of time. Most producing reservoirs can be regarded as
flowing under pseudo-steady state. ③ transient state refers to the condition under
which the reservoir pressures changes nonlinearly. This state typically occurs during
well shutdown or startup.
·72· Petroleum Production Technology

Fluids may flow through linear or radial pathways into the well. For instance, a
vertical production well is drilled, and fluids flow from all directions and finally
converge into the well. This pattern is named radial flow. For a horizontal well or a
fractured well, fluids flow linearly into the well, without convergence into the well.
This process is named linear flow.

4.2 Basic diffusivity equation

Under steady state, certain amount of fluid enters the porous medium and the same
amount of fluid leaves it. There is no accumulation of mass in the medium. However
under transient flow condition, the flow into the porous medium may not be the same
as the flow out. As a result, the amount of fluid in the medium changes with time. Now
consider the flow element in Fig. 4-1. The element is located at a distance from center
of well.

Fig. 4-1 Radial flow system

According to the material balance concept,


Mass flow rate entering volume element - Mass flow rate leaving volume element
= Rate of change of mass in the volume element
This can be expressed numerically as:
Chapter 4 Fluid Flow in Reservoir ·73·

⎛ ∂ ( qρ ) ⎞ ∂ρ
⎜ qρ |r + dr − qρ |r ⎟ = 2πrhφ dr (4-1)
⎝ ∂ r ⎠ ∂t

where,
q = flow rate of fluid, cm3/s;
ρ = density of fluid, g/cm3 ;
r = radius of reservoir, cm;
h = thickness of reservoir, cm;
φ = porosity of reservoir;
t = time, s.
Equation (4-1) can be simplified into:

∂ ( qρ ) ∂ρ
= 2πrhφ dr (4-2)
∂r ∂t

Recall Darcy’s law:


2πkhr ∂P
q= (4-3)
μ ∂r

where,
k = permeability of reservoir, D;
p = pressure, atm;
μ = viscosity of fluid, cP.
Substituting the flow rate q in Equation (4-2) with Equation (4-3) gives:

∂ ⎛ 2πkhr ∂P ⎞ ∂ρ
⎜ ρ ⎟ = 2πrhφ (4-4)
∂r ⎝ μ ∂r ⎠ ∂t

or,

1 ∂ ⎛ kρ ∂P ⎞ ∂ρ
⎜ r ⎟ =φ (4-5)
r ∂ r ⎝ μ ∂r ⎠ ∂t

Compressibility can be expressed as:

⎛m⎞
∂⎜ ⎟
ρ ⎝ ρ ⎠ 1 ∂ρ
c=− = (4-6)
m ∂P ρ ∂P
·74· Petroleum Production Technology

where,
c = compressibility, 1/atm①;
m = mass, g.
Differentiating with respect to time gives:
∂P ∂ ρ
cρ = (4-7)
∂t ∂t

Substituting equation (4-7) to equation (4-5) yields:

1 ∂ kρ ∂P ∂P
r = φ cρ (4-8)
r ∂ r μ ∂r ∂t

Equation (4-8) is the general partial differential equation that describes radial
flow of any single-phase fluid in porous media. Applying the chain rule to the equation
yields:

1⎛ ∂ k ∂P k ∂ρ ∂P kρ ∂P kρ ∂2 P ⎞ ∂P
⎜⎜ ρr + r + + r 2 ⎟⎟ = φ cρ (4-9)
r ⎝ ∂r μ ∂r μ ∂r ∂ r μ ∂r μ ∂r ⎠ ∂t

Compressibility can be differentiated with respect to radius:


∂ P ∂ρ
cρ = (4-10)
∂r ∂ r

Substituting equation (4-10) to equation (4-9) yields:

1 ⎡ ∂ ⎛ k ⎞ ∂P k kρ ∂P kρ ∂2 P ⎤
2
⎛ ∂P ⎞ ∂P
⎢ ⎜ ⎟ ρr + cρr ⎜ ⎟ + + r 2 ⎥ = φ cρ (4-11)
r ⎣⎢ ∂r ⎝ μ ⎠ ∂r μ ⎝ ∂r ⎠ μ ∂r μ ∂r ⎥⎦ ∂t

Assume rock permeability and fluid viscosity are constant. Moreover, the square
term is very small thus can be ignored. Equation (4-11) reduces to:

∂2 P 1 ∂P φ μc ∂P
+ = (4-12)
∂r 2
r ∂r k ∂t

Equation (4-12) can be more conveniently expressed as:


1 ∂ ⎛ ∂P ⎞ φ μc ∂P
⎜r ⎟= (4-13)
r ∂r ⎝ ∂r ⎠ k ∂t

————————
① 1 atm=1.01325×105Pa.
Chapter 4 Fluid Flow in Reservoir ·75·

Equation (4-13) is the radial diffusivity equation. The compressibility in the


equation is the total compressibility of the liquid-reservoir system that is expressed as:
ct = co So + cw S wc + cf (4-14)

where,
ct = total compressibility, 1/atm;
co = compressibility of oil, 1/atm;
cw = compressibility of water, 1/atm;
cf = compressibility of rock, 1/atm;
So = saturation of oil;
Swc = saturation of formation water.

4.3 Linear flow under steady state

Linear flow occurs in a reservoir when flow paths are parallel and the fluid flows in a
single direction. In addition, the cross sectional area to flow must be constant. A
common application of linear flow equations is the fluid flow into vertical hydraulic
fractures. This section describes steady-state liner flow of three fluid types:
incompressible fluid, slightly compressible fluid and compressible gas.

4.3.1 Steady-state linear flow of incompressible fluid


For an incompressible fluid flowing through the porous medium, the flow rate and
velocity remain constant at all locations in the medium. Separating variables in Darcy’s
law and integrating over the distance of flow gives:
q L k P2

A 0
dx = − ∫ dP
μ P1
(4-15)

where,
x = distance of flow, cm.
or,
kA( P1 − P2 )
q= (4-16)
μL

With field units, Equation(4-16) is expressed as:


·76· Petroleum Production Technology

0.001127kA( P1 − P2 )
q= (4-17)
μL

where,
q = flow rate, bbl/d;
k = absolute permeability, mD;
P = pressure, psi;
μ = viscosity, cp;
L = distance, ft;
A = cross-sectional area, ft2.

4.3.2 Steady-state linear flow of slightly compressible fluid


The expression below is valid for slightly compressible fluids:

V = Vref [1 + c( Pref − P) ] (4-18)

where,
V = volume of fluid, cm3;
Vref = volume of fluid under reference condition, cm3;
P = pressure, atm;
Pref = reference pressure, atm.
It can be modified in terms of flow rate:

q = qref [1 + c( Pref − P ) ] (4-19)

where,
q = fluid flow rate, cm3/s;
qref = fluid flow rate under reference condition, cm3/s.
Substituting the Equation(4-19) into Darcy’s equation gives:

q qref [1 + c( Pref − P ) ] k dP
= =− (4-20)
A A μ dx

where,
A = area that fluid flows across, cm2.
Separating the variables gives:
Chapter 4 Fluid Flow in Reservoir ·77·

qref L k P2 ⎡ dP ⎤
A ∫ 0 dx = − μ ∫ P1
⎢ ⎥
⎣1 + c( Pref − P) ⎦
(4-21)

where,
L = length of porous medium, cm;
P1 = pressure at entrance of porous medium, atm;
P2 = pressure at exit of porous medium, atm.
Rearranging the variables gives:

kA ⎡1 + c( Pref − P2 ) ⎤
qref = ln ⎢ ⎥ (4-22)
μcL ⎣ 1 + c( Pref − P1 ) ⎦

Selecting the upstream pressure (point 1) as the reference point gives:

kA
q1 = ln [1 + c( P1 − P2 )] (4-23)
μcL

The flow rate can be expressed in field units:


0.001127kA
q1 = ln [1 + c( P1 − P2 ) ] (4-24)
μcL

4.3.3 Steady-state linear flow of compressible gases


For gas flow in a linear system, the gas equation of state can be rearranged into:
PV
n= (4-25)
ZRT

Similarly, the equation is valid for standard conditions:


PscVsc
n= (4-26)
RTsc

where,
Psc = standard pressure, psia;
Tsc = standard temperature, ºR;
Vsc = volume of fluid under standard conditions, ft3.
Combining equation and gives:
·78· Petroleum Production Technology

PV PscVsc
= (4-27)
ZT Tsc

Equation (4-27) can be expressed in flow rate:

Pq Psc Qsc
= (4-28)
ZT Tsc

where,
q = gas flow rate at pressure p, ft3/d;
Qsc = fluid flow rate under standard conditions, ft3/d.
Z = gas compressibility factor;
Tsc = standard temperature, ºR;
Psc = standard pressure, atm.
Rearranging the variables gives:
Psc ZT
q= Qsc (4-29)
Tsc P

Combining Equation (4-29) with Darcy’s law gives:

q Psc ZT Qsc k dP
= =− (4-30)
A Tsc P A μ dx

Integrating over the distance and separating the variables gives:


Qsc PscT L P2 P
kTsc A ∫ 0 dx = ∫ P1 Zμg
dP (4-31)

where,
μg = gas viscosity, cP.
Assuming constant gas Z factor and gas viscosity between P1 and P2, and
integrating gives:

Tsc Ak ( P12 − P22 )


Qsc = (4-32)
2 PscTLZμg

If expressed in field units,


Chapter 4 Fluid Flow in Reservoir ·79·

0.003164Tsc Ak ( P12 − P22 )


Qsc = (4-33)
PscTLZμg

where,
Qsc = gas flow rate at standard conditions, scf/d;
k = permeability, mD;
T = temperature, ºR;
A = cross-sectional area, ft2;
L = total length of the linear system, ft.
Substituting values of standard pressure and temperature into Equation(4-33) gives:

0.111924 Ak ( P12 − P22 )


Qsc = (4-34)
TLZμg

It should be pointed out that gas Z factor and gas viscosity are actually strongly
affected by pressure. But they have been removed from the integral to simplify the
equations. Equation is valid for pressure lower than 2000 psi. Gas properties should be
evaluated at the average pressure defined below:

P12 + P22
P= (4-35)
2

Example: A gas flows linearly through a rock at 200ºF. Upstream and downstream
pressures are 1500 psia and 1300 psia, respectively. The gas has a specific gravity of
0.7. The length of the rock is 1000 ft, and its cross-sectional area is 1000 ft2. The rock
permeability is 100 mD. Calculate the gas flow rate.
The average pressure:

P12 + P22
P= = 1404 (psia)
2

Gas pseudo-critical pressure:

Ppc = 669.1

Gas pseudo-critical temperature:

Tpc = 389.4
·80· Petroleum Production Technology

Pseudo-reduced pressure:

Ppr = 1404/669.1 = 2.1

Pseudo-reduced temperature:

Tpr = (200+460)/389.4 = 1.7

Gas Z factor according to Fig. 1-6:

Z = 0.88

Gas viscosity with Lee method:

μg = 0.015 cP

Gas flow rate:

Qsc =
(
0.111924 Ak P12 − P22 ) = 0.111924 × 1000 × 100 × (1500 2
− 13002 )
TLZμg 660 × 1000 × 0.88 × 0.015
= 7.194 × 105 (scf / d)

4.4 Radial flow under steady state

For radial flow in a reservoir, fluids flow from all directions and finally converge into
the producing well. According to Darcy’s law, we understand fluids experience pressure
loss while flowing through the porous media. When the fluids finally reach the wellbore,
the fluid pressure is referred to as bottom-hole flowing pressure (BHP or Pwf).

4.4.1 Steady-state radial flow of incompressible fluid under


Fig. 4-1 schematically illustrates part of a radial flow system. It is assumes the
reservoir has a uniform thickness and a constant permeability. Because the fluid is
incompressible, the flow rate at any location is constant. Applying Darcy’s law to the
radial system gives:
q k dP
= (4-36)
Ar μ dr

At any point in the reservoir, the cross-sectional area at radius can be expressed as:
Chapter 4 Fluid Flow in Reservoir ·81·

Ar = 2πrh (4-37)

And Equation(4-36) becomes,

q q k dP
= = (4-38)
Ar 2πrh μ dr

The oil flow rate is often expressed in surface production rate (stb/day or stock
tank barrel per day), rather than reservoir units. Let Qo represent the flow rate in
stb/day, then

q = Bo Qo (4-39)

where,
Bo = oil formation volume factor,bbl/stb.
Equation (4-38) transforms into:
Qo Bo k dP
= (4-40)
2πrh μo dr

Assume two radii, r1 and r2, and the pressures at the two locations are P1 and P2.
Integrating Equation (4-40) yields:
r2 Qo dr P2 k
∫r1 2πh r
=∫
P1 μo Bo
dP (4-41)

Equation(4-41) can be simplified to:


Qo r2 dr k P2

2πh ∫ r1 r ∫P
= dP (4-42)
μo Bo 1

Performing the integration gives:


2πkh( P2 − P1 )
Qo = (4-43)
r
μo Bo ln 2
r1

Considering reservoir boundary and wellbore as the two radii, equation(4-43)


becomes
·82· Petroleum Production Technology

2πkh( Pe − Pwf )
Qo = (4-44)
r
μo Bo ln e
rw

The flow rate can be expressed in field units:


0.00708kh( Pe − Pwf )
Qo = (4-45)
r
μo Bo ln e
rw

where,
Qo = oil flow rate, stb/d;
Pe = external pressure, psi;
Pwf = bottom–hole flowing pressure, psi;
k = permeability, mD;
μo = oil viscosity, cP;
Bo = oil formation volume factor, bbl/stb;
h = thickness, ft;
re = external or drainage radius, ft;
rw = wellbore radius, ft.
Equation(4-45) can be rearranged to find the pressure at any radius:
Qo Bo μo r
P − Pwf = ln (4-46)
0.00708kh rw

For steady state, the reservoir average pressure is located at 61% of the drainage
radius (re). Equation(4-46) can be expressed as:
0.00708kh( Pr − Pwf )
Qo = (4-47)
0.61re
μo Bo ln
rw

or,
0.00708kh( Pr − Pwf )
Qo = (4-48)
⎛ r ⎞
μo Bo ⎜ ln e − 0.5 ⎟
⎝ rw ⎠

Example: A well is producing from a reservoir under steady state. The following
information is available. Find the pressure at 5 ft, 10 ft, 20 ft, 50 ft, 100 ft, 200 ft, 400 ft,
Chapter 4 Fluid Flow in Reservoir ·83·

600 ft, 800 ft and 1000 ft away from center of well.


Oil production rate = 1000 stb/d
Well flowing pressure = 1000 psia
Reservoir permeability = 100 mD
Reservoir thickness = 100 ft
Reservoir radius = 1000 ft
Wellbore radius = 0.35 ft
Oil viscosity in situ = 10 cP
Oil formation volume factor = 1.1
The pressures at various locations can be calculated with Eq.(4-45). The results
are summarized in Tab. 4-1.
Tab. 4-1 Pressure profile in reservoir
Radius/ft Pressure/psia
1 1163
3 1334
5 1413
10 1521
20 1629
50 1771
100 1879
200 1986
300 2049
400 2094
500 2129
700 2181
1000 2236

The pressure profile is further plotted in Fig. 4-2. While fluid moves from
reservoir boundary to 400 ft, the pressure declines almost linearly from 2236 psi to
2094 psi. Afterwards, pressure quickly drops to 1000 psi when the fluid reaches the
wellbore. The pressure declines significantly for all fluids to converge into the well. In
conclusion, most pressure loss occurs in the near-wellbore region for radial flow
system.
·84· Petroleum Production Technology

Fig. 4-2 Reservoir pressure profile

4.4.2 Steady-state radial flow of compressible gases


For radial gas flow, Darcy’s law takes the form:
2πrhk dP
qgr = (4-49)
μg dr

where,
qgr = gas flow rate at radius r, bbl/d;
r = radial distance, ft;
h = zone thickness, ft;
μg = gas viscosity, cP;
P = pressure, psi.
Gas flow rate can be converted to that under standard conditions by gas equation
of state:
qgr P Qg Psc
= (4-50)
ZRT RTsc

or,
Psc ZT
qgr = Qg (4-51)
Tsc P

where,
Chapter 4 Fluid Flow in Reservoir ·85·

Qg = gas flow rate at standard conditions, scf/d;


qgr = real gas flow rate at a certain radius, ft3/d.
Combining Equation (4-49) and Equation (4-51) yields:

Psc ZT 2πrhk dP
Qg = (4-52)
Tsc P μg dr

Rearrange Equation(4-52), thus

TQg dr T 2P
= π sc dP (4-53)
kh r Psc μg Z

Integrating between wellbore and any point in the reservoir gives:

r TQg dr T P 2P
∫ rw kh r = π Pscsc ∫Pwf μg Z
dP (4-54)

Assume temperature, permeability and reservoir thickness are constant, thus

TQg ⎛ r ⎞ Tsc P 2P
ln ⎜
kh ⎝ rw
⎟=π
⎠ Psc ∫P wf μg Z
dP (4-55)

The integral term in Equation (4-55) can be expressed as:


P 2P P 2P Pwf 2 P
∫P
wf μg Z
dP = ∫
0 μg Z
dP − ∫
0 μg Z
dP (4-56)

Equation (4-55) transforms into:

TQg ⎛ r ⎞ Tsc ⎛ P 2 P Pwf 2 P ⎞


ln ⎜ ⎟=π ⎜∫ dP − ∫ dP ⎟ (4-57)
kh ⎝ rw ⎠ Psc ⎜⎝ 0 μg Z 0 μg Z ⎟

The integral is named real gas potential, and it is represented with ψ .


P 2P
ψ =∫ dP (4-58)
0 μg Z

Thus Equation(4-57) becomes


·86· Petroleum Production Technology

TQg r T
ln = π sc (ψ − ψ w ) (4-59)
kh rw Psc

The gas flow rate can be expressed as:


Tsc πkh(ψ − ψ w )
Qg = (4-60)
Psc r
Tln
rw

Considering reservoir boundary and field units,


0.703kh(ψ e −ψ w )
Qg = (4-61)
r
T ln e
rw

4.5 Fluid flow under pseudo-steady state

Under pseudo-steady state, the reservoir has been producing for a sufficiently long time,
so that the effect of out boundary has been felt. No fluid flows into the reservoir from
the reservoir outer boundary. Thus,
∂P
= 0 at r = re
∂r

where,
re = radius of reservoir.
Moreover, if the well produces at a constant rate, then the reservoir pressure
∂P
declines linearly. ≈ constant, for all locations and time
∂t
According to the definition of compressibility:
dP dV
cV =− = −q (4-62)
dt dt

or,
dP q
=− (4-63)
dt cV
Chapter 4 Fluid Flow in Reservoir ·87·

Substituting the volume component yields:


dP q
=− (4-64)
dt cπre2 hφ

Substituted in the radial diffusivity Equation (4-13) gives:

1 ∂ ∂P qμ
r =− 2 (4-65)
r ∂r ∂r πre kh

Integration yields:

∂P qμr 2
r =− + C1 (4-66)
∂r 2πre2 kh

Applying the boundary condition to find:



C1 = (4-67)
2πkh

Thus Equation (4-66) becomes:

∂P qμ ⎛ 1 r ⎞
= ⎜ − ⎟ (4-68)
∂r 2πkh ⎜⎝ r re2 ⎟⎠

Integrating again gives:


r
qμ ⎡ r2 ⎤
[ P] P
Pwf
= ⎢ln r − 2 ⎥ (4-69)
2πkh ⎢⎣ 2re ⎥⎦ rw

or,

qμ ⎛ r r2 ⎞
P − Pwf = ⎜⎜ ln − 2 ⎟⎟ (4-70)
2πkh ⎝ rw 2re ⎠

When r = re,

qμ ⎛ re 1 ⎞
Pe − Pwf = ⎜ ln − ⎟ (4-71)
2πkh ⎝ rw 2 ⎠

It is more convenient to express equation with average reservoir pressure. The


·88· Petroleum Production Technology

volume averaged pressure in the radial cell is expressed as:


re
∫ r PdV
Pr = w
re
(4-72)
∫r w
dV

And the volume component is given:


dV = 2πrhφ dr (4-73)

Equation (4-72) transforms into:


re
∫ r P2πrhφ dr
Pr = w
(4-74)
π(re2 − rw2 )hφ

or,
2 re
Pr =
(re2 − rw2 )
∫r
w
Prdr (4-75)

The term rw2 is small and ignored. Substituting Equation (4-75) to Equation (4-70)
yields:

2 qμ re ⎛ r r2 ⎞
P − Pwf = ∫ r ⎜ ln − ⎟ dr
re2 2πkh rw ⎜⎝ rw 2re2 ⎟⎠
(4-76)

Integration by parts gives:


r r r
re r ⎡ r2 r ⎤
e
re 1 r 2 ⎡ r2 r ⎤
e
⎡ r2 ⎤ e
∫ rw rw
r ln dr = ⎢ ln ⎥ − ∫ rw r 2 d r = ⎢ ln ⎥ − ⎢ ⎥
⎣ 2 rw ⎦ rw ⎣ 2 rw ⎦ rw ⎣ 4 ⎦ rw
re2 re re2
≈ ln −
2 rw 4

On the other hand,


r
re r3 ⎡ r 4 ⎤ e re2
∫r 2re2
dr = ⎢ 2⎥ ≈
8
w ⎣⎢ 8re ⎦⎥ r w
Chapter 4 Fluid Flow in Reservoir ·89·

Combing the two results gives:

qμ ⎛ re 3 ⎞
P − Pwf = ⎜ ln − ⎟ (4-77)
2πkh ⎝ rw 4 ⎠
Chapter 5 Drilling and Production Facilities
Wells are drilled to reach the oil and gas in the reservoir. Drilling is a very expensive
operation that involves many skilled staff and service providers. Drilling costs a few
million US dollars for a well on land, to more than 100 million US dollars for a well
offshore. This chapter presents the surface facilities involved in drilling and production
operations.

5.1 Drilling rigs

For drilling on land, the drilling rig is transported to the designated site, assembled and
erected, as seen in Fig. 5-1. A traditional rotary drilling rig employs hoisting equipment,
mud circulation equipment, and rotary equipment. These facilities will be introduced later.

Fig. 5-1 Onshore drilling rig

For offshore drilling, various types of rig are available for different water depths
(Fig. 5-2). Swamp barges operate in very shallow water (less than 6m). They can be
towed onto location then ballasted so that they sit on bottom. The drilling unit is
mounted onto the barge. This type of unit is used in the swamp areas of Nigeria,
Venezuela and US Gulf Coast.
Drilling jackets are small steel platform structures which are used in areas of
shallow and calm water. A number of wells may be drilled from one jacket. If a jacket
Chapter 5 Drilling and Production Facilities ·91·

is too small to accommodate a drilling operation, a jack-up rig is usually cantilevered


over the jacket and the operation carried out from there.

Fig. 5-2 Offshore drilling structures for different water depths

Jack-up rigs are operational in water depths up to about 150m and as shallow as
50m. Globally, they are the most common rig type, used for a wide range of
environments and all types of wells. Jack-up rigs are either towed to the drilling
location or they are equipped with a propulsion system. The three or four legs of the rig
are lowered onto the seabed. After some penetration the rig will lift itself to a
determined operating height above the sea level. All drilling and supporting equipment
are integrated into the overall structure (Fig. 5-3).

Fig. 5-3 A jack-up rig


·92· Petroleum Production Technology

Semi-submersibles are used for drilling in water depths too great for a jack-up. A
semi-submersible rig is a movable offshore vessel consisting of a large deck area built
on columns of steel (Fig. 5-4). Attached to these heavy-duty columns are at least two
barge-shaped hulls called pontoons. Before operation commences on a specified
location, these pontoons are partially filled with water and submersed in approximately
15m of water to give stability. A large-diameter steel pipe named riser is connected to
the seabed and serves as a conduit for the drill string. The blowout preventer is also
located at the seabed. A combination of several anchors and dynamic positioning
equipment maintaining its position. Relocation of the semi-submersible vessel is made
possible by the tug boats and/or propulsion machinery. Heavy-duty semi-submersibles,
for example the Deepwater Horizon which is rated 15,000 psi, can handle high
reservoir pressures and operate in the most severe conditions in water depths down to
3000m.

Fig. 5-4 A semi-submersible rig

Drill ships are used for deep and very deep drilling (Fig. 5-5). They can be less
stable in rough seas than semi-submersibles. However, modern drill ships such as
Discoverer Enterprise can remain stable in 100-knot winds using powerful thrusters
controlled by a dynamic positioning system. The thrusters counter the forces of
currents, wind and waves to keep the vessel exactly on target, averaging less than 2m
off its mark. Heavy-duty drill ships are capable of operating in water depths up to
3000m without an anchor.
Chapter 5 Drilling and Production Facilities ·93·

Fig. 5-5 A drill ship

5.2 Drilling systems

Common drilling systems include rotary drilling system and top drive system. Rotary
drilling system is often employed in low-cost onshore areas, while the top drive system
is now dominant in offshore drilling. Even though these two systems rely on different
drive mechanisms, they both employ a drill string, drill bit, drilling fluid, and blowout
preventer.

5.2.1 Rotary drive and top drive


A traditional rotary drilling system is illustrated in Fig. 5-6. Torques are produced by
diesel engines, then transmitted to the rotary table through gears underneath. The rotary
table spins the kelly as well as the drill string attached. During drilling, the drill string
is raised or lowered by the hoisting system that consists of draw-works, wire cable,
crown block, travelling block and hook.
Instead of having a rotary table and kelly in the rig floor, the top drive is mounted
on guide rails and moves up and down inside the derrick (Fig. 5-7 ). This allows drilling
in segments of 90 ft of pre-assembled pipe, significantly reducing connection time. The
more continuous drilling process also results in better hole conditions and faster
penetration rates. The latest rigs allow running of 120 ft sections and are equipped with
two derricks: One drilling the well and the other concurrently pre-assembling drill strings.
·94· Petroleum Production Technology

Fig. 5-6 A typical rotary drilling system

Fig. 5-7 Top drive system


Chapter 5 Drilling and Production Facilities ·95·

5.2.2 Drill string


The drill string is made up of drill pipes, drill collars, drill bit and other accessories
(Fig. 5-8). Drill pipes are normal steel pipes. Drill collars are thick-walled pipes that
keep the drilling string in tension and add weight to bit. Drill bit is the tool that breaks
rock into pieces while it rotates with the drill string. Two types of bit are available in
market, roller cone bit (Fig. 5-9 ) and PDC bit (polycrystalline diamond compact bit
in Fig. 5-10).
Drilling technology has advanced significantly in recent years. Nowadays, most of
the wells being drilled are horizontal wells, directional wells (Fig. 5-11) and
multilateral wells (Fig. 5-12). These complex wells require more advanced and precise
drilling techniques, but they provide much higher productivity than traditional vertical
wells. Drilling operation is briefly introduced in this chapter.

Fig. 5-8 Typical components on a drill string Fig. 5-9 Roller cone bit
·96· Petroleum Production Technology

Fig. 5-10 PDC bit

Fig. 5-11 Common well trajectories

Fig. 5-12 Multilateral wells


Chapter 5 Drilling and Production Facilities ·97·

5.2.3 Mud circulation


While drilling progresses, drilling mud is circulated in the well at all times. The
purpose of mud circulation is to cool and lubricate the string. Drilling mud also
controls formation pressure and carries the broken rock pieces (namely cuttings) to
surface. Water-based mud is made by mixing clay, weighting materials and chemicals
into water. Oil-based mud is also available in the market. Oil-based mud provides
better lubrication and well-bore stability. But it is more expensive, flammable and toxic
to environment.
A mud pump provides the energy for mud circulation. Swivel allows mud to enter
drilling string and contains bearings for rotation of the drill string. Mud flows through
the drill string and the annulus. Blowout preventers are in place to shut down mud flow
in case of an emergency. After mud returns to surface, it is cleaned sequentially by
shale shaker, desander, desilter and made ready for another circulation. Mud circulation
system is illustrated in Fig. 5-13 .

Fig. 5-13 Mud circulation system

5.2.4 Blowout preventer


An important safety feature employed by every modern rig is the blowout preventer
(BOP). Drilling mud provides a hydrostatic head of fluid to counterbalance the pore
·98· Petroleum Production Technology

pressure of fluids in permeable formations. However, formation fluids may enter the
wellbore, pushing mud out of the hole. This phenomenon is named well kick. If left
uncontrolled, this can lead to a blowout, a situation where formation fluids flow to the
surface in an uncontrolled manner.
The BOP contains a series of powerful sealing elements designed to close off the
annular space between the pipe and the hole, through which the mud normally returns
to the surface. By closing off this route, the well is shut in and the mud and/or
formation fluids must flow through a controllable choke or adjustable valve. This
choke allows the drilling crew to control the pressure that reaches the surface.
The crew will follow necessary steps to kill the well and restore a balanced system.
Fig. 5-14 shows a typical BOP stack. The annular preventer has a rubber sealing
element that is hydraulically inflated to fit tightly around any size of pipe in the hole.
Ram type preventers either grip the pipe with rubber-lined steel pipe rams, block the
hole with blind rams when no pipe is in place or cut the pipe with powerful hydraulic
shear rams to seal off the hole. BOP is opened and closed by hydraulic fluid stored
under a pressure of 3000 psi in an accumulator.

Fig. 5-14 BOP stack

5.2.5 Casing program


When a certain depth is reached, casings are inserted into the well and cemented.
Casings are steel pipes in nature. After the top section is drilled, the conductor casing is
inserted and cemented in place. Subsequently, surface casing, intermediate casing and
production casing are installed and cemented (Fig. 5-15). Sometimes, a production
liner is run instead of production casing. Liner is connected to the intermediate casing
Chapter 5 Drilling and Production Facilities ·99·

by a liner hanger. Liner does not run to surface, thus it is a cheaper option.
The outer diameter of conductor pipe is often between 20 and 30 in. When drilling
the subsequent section, a smaller bit is used to fit into the previous stage of casing. It is
obvious the subsequent stage of casing is also smaller to be able to go through the
previous casing. The diameter of production casing or liner is usually 8 to 10 in only.

Fig. 5-15 Typical casing program

Casings serve the following important functions. ① To support weak formations.


Without casing, the well may cave in and well is buried. ② To prevent contamination
of fresh water zone located near surface. ③ To provide connection for blowout
preventer and wellhead equipment. ④ To isolate production zones. A well with
casings and cement is referred to as cased well or cased hole. In fact, other completion
options are also available. Details of well completion will be discussed in Chapter 6.
A simple production system is presented in Fig. 5-16. In this scenario, casings are
cemented in the well. Tubings, wellhead and surface separators are installed. The
production zone is perforated so that reservoir fluids can flow into the well. Fluids flow
upwards through several thousand meters of tubing to reach wellhead. Then the fluid
stream enters separators and breaks into water, oil, gas and solids. Produced water is
treated then discarded or re-injected into the reservoir. The produced oil and gas are
transported to refineries for further processing. Well completion methods will be introduced
in Chapter 6. Separation and treatment of fluids will be discussed in Chapter 13.
·100· Petroleum Production Technology

Fig. 5-16 Overview of petroleum production system

5.3 Offshore production structures

Several tybepes of offshore structures are available as offshore production platforms.


They can be categorized as fixed structures and floating structures. Fixed structures
include steel jackets and concrete structures. Floating structures include tension leg
platform, FPSO (floating production, storage and offloading vessel), semisubmersible
and SPAR platform. The choice of suitable offshore production structure is mainly
influenced by the water depth, as illustrated in Fig. 5-17.

Fig. 5-17 Offshore structures at a glance


Chapter 5 Drilling and Production Facilities ·101·

Steel piled jackets are the most common type of platform and are employed in a
wide range of sea conditions, from the calm South China Sea to the hostile Northern
North Sea. Steel jackets are used in water depths of up to 150m and may support
production facilities a further 50m above mean sea level. In deepwater, all the process
and support facilities are normally supported on a single jacket (Fig. 5-18), but in
shallow seas it may be cheaper and safer to support drilling, production and
accommodation modules on different jackets.

Fig. 5-18 A steel jacket

Steel jackets are constructed from welded steel pipes. The jacket is fabricated
onshore and then floated out horizontally on a barge and set upright on location (Fig.
5-19). Once in position a jacket is pinned to the seafloor with steel piles. Prefabricated

Fig. 5-19 Transportation of a steel jacket on a barge


·102· Petroleum Production Technology

units or modules containing processing equipment, drilling and other equipment are
installed by lift barges on to the top of the jacket, and the whole assembly is connected
and tested by commissioning teams. Steel jackets can weigh 20000 t or more and
support a similar weight of equipment.
Concrete or steel gravity-based structures can be deployed in similar water depths
to steel jacket platforms. A concrete platform is shown in Fig. 5-20. Gravity-based
platforms rely on weight to secure them to the seabed, which eliminates the need for
piling in hard seabed. Concrete gravity-based structures are built with huge ballast
tanks surrounding hollow concrete legs (Fig. 5-20). They can be floated into position
without a barge and are sunk once on site by flooding the ballast tanks. For example,
the Mobil Hibernia Platform offshore Canada weighs around 450000 t and is designed
and constructed to resist iceberg impact.

Fig. 5-20 A concrete platform

The legs of the platform can be used as settling tanks or temporary storage
facilities for crude oil (Fig. 5-21) where oil is exported via tankers, or to allow
production to continue in the event of a pipeline shutdown. The Brent D Platform in the
North Sea weighs more than 200000 t and can store over a million barrels of oil.
Floating production systems are becoming much more common as a means of
developing smaller fields which cannot support the cost of a permanent platform and
for deepwater development. Ships and semi-submersible rigs have been converted or
custom built to support production facilities which can be moved from field to field as
reserves are depleted.
Chapter 5 Drilling and Production Facilities ·103·

Fig. 5-21 A concrete platform with ballast tanks

Tension leg platforms (TLP) are used in deepwater where fixed platforms would
be both vulnerable to bending stresses and very expensive to construct. TLP is tethered
to the seabed by jointed legs kept in tension like a semi-submersible rig (Fig. 5-22).
Tension is maintained by pulling the floating platform down into the sea below its
normal displacement level. The legs are secured to a template or anchor points installed
on the seabed.
·104· Petroleum Production Technology

Fig. 5-22 A tension leg platform

The latest generation of floating production systems have the capacity to deal with
much more variable production streams and additionally store and offload crude, and
hence are referred to as FPSO (floating production, storage and offloading vessel).
The newer vessels can provide all services which are available on integrated platforms,
in particular three-phase separation, gas lift, water treatment and injection (Fig. 5-23,
Fig. 5-24).

Fig. 5-23 A floating production storage and offloading vessel(FPSO)


Chapter 5 Drilling and Production Facilities ·105·

Fig. 5-24 FPSO with tanker ship and subsea production system

Ship-shaped FPSO vessels have the ability to rotate in the direction of wind or
current. This requires complex mooring systems and the connections with the
wellheads must be able to accommodate the movement. The typical process capability
for FPSO is around 100000 barrels per day, with storage capacity up to 800000 bbl. In
the recent deepwater developments in West Africa, some FPSO now double this
capacity.
Semi-submersible vessels have been used for many years as floating production
systems, examples include Argyll, Buchan, Balmoral and Veslefrikk in the North Sea,
and more recently Mad Dog and Thunder Horse in more than 1800 m of water in the
gulf of Mexico. In Brazil, the Campos and Roncador fields employ semi-subs in 930
and 1360 m of water, respectively.
A semi-sub may either be a new build, for example Balmoral, or a converted
drilling rig, for example Argyll and Buchan (Fig. 5-25). Conversion offers the
potential for short lead times but may limit topsides weight capacity compared to
recent new builds like Thunder Horse. New semi-subs can be built very large and
designed to accommodate very large deck loads and support heavy steel risers.
SPAR platforms were first employed as a concept by Shell, when it was used as a
storage facility for the Brent field in the North Sea. It had no production facilities but
was installed simply for storage and offshore loading. More recently, SPAR structures
have incorporated drilling, production, storage and offshore loading facilities as an
integrated development option.
·106· Petroleum Production Technology

Fig. 5-25 A semi-submersible platform

Examples of SPAR developments include Genesis, Neptune, Hoover and the


Devils Tower Fields in the Gulf of Mexico, the latter of which was installed in 1710 m
water depth. A SPAR platform is shown in Fig. 5-26. It is usually transported on a
barge (Fig. 5-27). SPAR platforms are more common in the Gulf of Mexico than
anywhere else, as until 2006 the U.S. authorities would not license FPSO in the gulf
waters due to environmental concerns associated with such systems.

Fig. 5-26 SPAR platform in operation


Chapter 5 Drilling and Production Facilities ·107·

Fig. 5-27 Transportation of SPAR on a barge

SPAR can support dry trees and allow conventional access to development wells.
It is also relatively easy to tie in additional wells to a SPAR during its lifetime.
However, the size of buoyancy compartments required to support steel risers limits the
water depth in which this concept can be employed.
Chapter 6 Well Completion
After a well has been drilled, multiple options are available to complete the well. Well
completion is the process of making a well ready for production or injection. This
principally involves preparing the bottom of the hole to the required specifications,
running in the production tubing and its associated down-hole tools as well as
perforation and stimulation as required. Completions include lower completion (also
named reservoir completion) and upper completion. Lower completion refers to the
connection between reservoir and well. The upper completion refers to the conduit for
produced fluids to reach surface.

6.1 Lower completion

Several options are available for reservoir/lower completion, as shown in Fig. 6-1. For
open-hole completion, the deepest casing is set on top of the producing interval, and the
producing zone is fully open. For cased hole, casing is cemented through the producing
interval. Communication with the reservoir is established by perforating the casing.

6.1.1 Open-hole completion


Open-hole completion refers to a range of completion methods where no casing or liner
is cemented in place across the production zone. The advantages of open-hole
completion include: ① higher productivity; ② no perforating costs and related
damage; ③ larger hole diameter. Its limitations are: ① difficult to control excessive
gas or water production; ② selective fracturing or acidizing is difficult; ③ may
require frequent cleanout.
In competent formations, the zone might be left entirely open, but some sort of
sand-control and/or flow-control tools are usually incorporated. For example, predrilled
liner, slotted liner, or sand screens are often installed in an open hole well. Predrilled
liner is prepared with multiple small drilled holes (Fig. 6-2), then set across the
production zone to provide wellbore stability.
Chapter 6 Well Completion ·109·

Fig. 6-1 Open-hole (a), cased-hole (b), and liner completion (c)

Fig. 6-2 Predrilled liners


·110· Petroleum Production Technology

Predrilled liners are often combined with open hole packers, such as swelling
packers, mechanical packers or external casing packers, to provide zonal segregation
and isolation. It is now quite common to employ a combination of predrilled liner and
swell packers to initially isolate unwanted water or gas zones. Multiple sliding sleeves
can also be used in conjunction with open hole packers to provide considerable
flexibility in zonal flow control for the life of the wellbore. Packers will be introduced
later in this chapter.
Slotted liners (Fig. 6-3) can be an alternative to predrilled liner. A slotted liner is
machined with multiple longitudinal slots (for example 2mm×50mm) spread across
the length and circumference of each joint. Traditionally, the slots are cut with small
rotary saws. Recent advance in laser cutting technology allows slotting be done much
cheaper to much smaller slot widths.

Fig. 6-3 Slotted liners

Several patterns for slots layout can be found in market, as seen in Fig. 6-4. The
single-slot staggered, longitudinal pattern is generally preferred, because the strength of
the original pipe is preserved. The staggered pattern also gives a more uniform
distribution of slots over the surface area of the pipe. The single-slot staggered pattern
is slotted with an even number of rows around the pipe with a typical 6 in longitudinal
spacing of slot rows.
Chapter 6 Well Completion ·111·

Fig. 6-4 Patterns for slots

The slots can be straight or keystone shaped, as illustrated in Fig. 6-5. The
keystone slot is narrower on the outside surface of the pipe than on the inside. Slots
formed in this way have an inverted “V” cross-sectional area and are less prone to
plugging, because any particle passing through the slot at the outside diameter of the
pipe will continue to flow through, rather than lodging within the slot.

Fig. 6-5 Straight slot and keystone slot

Predrilled liners and slotted liners are generally not very effective at sand control.
It is difficult to make the holes or slots small enough to stop sand particles. Of course,
laser can cut very tiny slots, however very small slots are susceptible to plugging.
Generally speaking, predrilled liners are preferred over slotted liner, because they are
stronger than slotted liners, especially under collapse pressure and installation torque.
The small holes on predrilled liners ensure much larger area for fluid flow, thus
reducing the chocking effect on production rate.
·112· Petroleum Production Technology

6.1.2 Cased-hole completion


As described earlier, cased-hole completion involves running casing or liner down
through the production zone, and cementing it in place. Connection between the
well-bore and the formation is made by perforating. Sand control devices can be
installed in cased wells. Because perforation intervals can be precisely positioned,
cased-hole completion leads to good control of fluid flow, although it relies on the
quality of the cement to prevent fluid flow behind the liner or casing. Other advantages
of cased-hole completion include: ① gas or water production can be more readily
controlled. ② selective stimulation can be carried out. Its limitations include:
③ perforating is expensive and may result in formation damage. ④ logging is crucial
to recognize producing intervals.
Multiple logging tools are lowered into the well before the production casing is
cemented. These tools normally include permeable zone logging tools, resistivity
logging tools and porosity logging tools. Logging can also be done while drilling. This
operation is referred to as logging while drilling (LWD). Multiple logging tools are
attached to the drill string and logging is conducted while drilling proceeds.
① Permeable zone logs, such as spontaneous potential log and gamma ray log, record
the natural charges and gamma ray emissions to locate the reservoirs. ② Resistivity
log tools send electric currents into the rocks, the rock zones that contain oil and gas
show high resistivity because oil and gas do not transmit electricity. ③ Porosity logs
include density log, neutron log and sonic log. Density log tool emits gamma ray into
formations and detect the returned gamma ray, and determines rock density and
porosity accordingly. Neutron log tool emits fast neutrons into formations and detects
the returned neutron, thus determines rock porosity. Sonic log tool sends sound waves
into rock formations and records the sound travel time to the receiver, thus determining
porosity by the speed of sound in the rock.
Sand control is often necessary for both open hole and cased wells. There are
many variants of sand control methods, the three popular choices being stand-alone
screens, expandable screens and gravel packs. Sand control methods will be introduced
in chapter 11.

6.1.3 Choice of lower completion


Many engineers do not routinely consider an open-hole completion when confronted
Chapter 6 Well Completion ·113·

with completion design. This is true probably because cased-hole completions are so
widely accepted, and engineers are not familiar with selection criteria and procedures.
Open-hole completions can provide excellent, high-productivity completions, but
they must be applied under the right reservoir conditions. Open-hole completions can
avoid the difficulties and concerns of perforating, and reduce the gravel-placement
operations to the relatively simple task of packing the screen and open-hole annulus.
Because open-hole wells have no perforation tunnels, formation fluids can converge
toward and through the gravel pack radically from all directions, eliminating the high
pressure drop due to linear flow through perforation tunnels. The reduced pressure drop
through an open-hole gravel pack virtually guarantees that it will be more productive
than a cased-hole gravel pack in the same formation, provided they are executed
properly.
Open hole completions have seen significant uptake in recent years, and many
configurations have been developed to address specific reservoir challenges. Many
recent developments have boosted the success of open-hole completions. In practice,
open-hole completions are more common in carbonate fields. They are also very
popular for horizontal wells, where cementing is more expensive and more difficult
technically. It is basically the same as the vertical open-hole completion, but an
open-hole horizontal well yields significantly larger contact with the reservoir, thus
increasing the production or injection rate of the well.

6.1.4 Cementing a cased well


As part of the process of preparing a well for further drilling, production or abandonment,
cementing a well is the procedure of preparing and pumping cement into place. After
casing is run into a drilled well, cement is injected into the well via pumps, displacing
the drilling mud still inside the well, and replacing it with cement. The cement slurry
flows to the bottom of the wellbore through casing. From there it fills in the annulus
between the casing and the actual wellbore, and hardens. This creates a seal and
permanently positions the casing in place.
The most important function of cement is to achieve long-term zonal isolation.
Besides, cement also has the following functions: ① support the vertical and radial
loads applied to the casing; ② isolate porous formations from the producing zone
formations; ③ exclude unwanted subsurface fluids from the producing interval;
④ protect casing from corrosion; ⑤ cementing is used to plug and abandon a well.
·114· Petroleum Production Technology

Cement was first discovered by an English brick layer named Joseph Aspdin in 1824.
He called it Portland cement for the reason that the cement resembled the limestone
found in Portland. Portland cement contains 60%~70% lime (CaO), 20%~25%
silica (SiO2), 5%~10% alumina (Al2O3), and 2%~3% ferric oxide (Fe2O3).
Cementing service companies stock various types of cement and have special
transportation equipment to handle this material in bulk. Cement storage and handling
equipment is moved out to the rig, making it possible to mix large quantities of cement
at the site. The cementing crew mixes the dry cement with water, using a device called
a jet-mixing hopper. The dry cement is gradually added to the hopper, and a jet of water
thoroughly mixes with the cement to make a slurry. The amount of time it takes cement
to harden is called thickening time. For setting wells at deep depths, under high
temperature or pressure (for example steam injection wells and geothermal wells), as
well as in corrosive environments, special cements can be employed. The American
Petroleum Institute (API) has classified 9 types of cement as follows.
Class A is intended for 0 to 6000 ft (1830 m) depth, when special properties are
not required. It is available only in ordinary type. Class B is intended for 0 to 6000 ft
depth, when conditions require moderate to high sulfate resistance. It is available in
moderate and high sulfate resistant types. Class C is intended for 0 to 6000 ft depth,
when conditions require high early strength. It is available in ordinary, moderate and
high sulfate resistant types.
Class D is intended for 6000 to 10000 ft (1830 to 3050 m) depth under
moderately high temperature and pressure. It is available in moderate and high sulfate
resistant types. Class E is intended for 10000 to 14000 ft (3050 to 4270 m) depth
under conditions of high temperature and pressure. It is available in moderate and high
sulfate resistant types. Class F is intended for 10000 to 16000 ft (3050 to 4880 m)
depth under conditions of extremely high temperature and pressure. It is available in
moderate and high sulfate resistant types.
Class G is intended for 0 to 8000 ft (2400 m) depth. It can be used with
accelerators and retarders to cover a wide range of well depths and temperatures. Class
G is available in moderate and high sulfate resistant types. Class H is intended for 0 to
8000 ft (2400 m) depth, can be used with accelerators and retarders to cover a wide
range of well depths and temperature. It is available only in moderate sulfate resistant
type. Class G and H cements are manufactured to more rigorous chemical and physical
specifications, resulting in a more uniform product. They contain no accelerators,
retarders, or viscosity-control agents. They are compatible with these chemicals for use
Chapter 6 Well Completion ·115·

over the complete range of conditions. In practice, they can be used in any cementing
conditions with proper additives. Worldwide, class G cement is replacing most other
API cements.
Class J is intended for 12000 to 16000 ft (3660 to 4880 m) depth, can be used
under conditions of extremely high temperature and pressure, or can be mixed with
accelerators and retarders to cover a range of well depth and temperatures.
In preparing a well for cementing, it is important to establish the amount of
cement required for the job. This is done by measuring the diameter of the borehole
along its depth, using a caliper log. Utilizing both mechanical and sonic means, caliper
logs measure the diameter of the well at numerous locations simultaneously in order to
accommodate for irregularities in the wellbore diameter and determine the volume of
the open hole.
Additionally, the required physical properties of the cement are essential before
commencing cementing operations. The proper cement properties are also determined,
including the density and viscosity of the material, before actually pumping the cement
into the hole. The thickening time of a cement slurry is an important factor in slurry
design. The primary factoring affecting thickening time is temperature. The cement sets
faster at higher temperatures, as seen in Tab. 6-1. High pressure and loss of water also
shorten thickening time.
Tab. 6-1 Effect of temperature on thickening time
Circulating temperature Thickening time for Thickening time for
/ºF class G cement/min class H cement/min
103 150 200
113 130 117
125 104 100

Common cement additives include accelerator, retarder, extender, antifoaming


agent, and weight materials. Accelerators shorten the setting time required for the slurry.
Retarders do the opposite and make the cement setting time longer. Extenders can
expand the cement in an effort to reduce the cost of cementing. Antifoam additives can
be added to prevent foaming within the well. Bridging materials are required to plug
lost circulation zones.
When cementing shallow wells, it may be necessary to accelerate the cement
hydration to reduce waiting time. Common cement accelerators are calcium chloride,
sodium chloride, gypsum, and sodium silicate. Calcium chloride is the most widely-
·116· Petroleum Production Technology

used accelerator and available in regular 77% grade. A dosage of 2% regular grade is
often used. Above 3%, little additional acceleration is gained. A dosage above 4% is
detrimental to long-term cement strength. Sodium chloride in low concentration
accelerates cement setting, producing optimum acceleration at 2 to 2.5wt%.
Increasing well depths and higher formation temperatures led to the development
of retarders that extend thickening time. A common cement retarder is calcium
lignosulfonate. It can function at temperatures up to 290ºF. It also reduces cement
viscosity when necessary.
Lightweight and heavyweight materials are added to decrease or increase the
density of the cement. Lightweight solids that reduce slurry density include bentonite
and solid hydrocarbons. Heavyweight solids that increase slurry density include barite
and hematite.
The cementing process is depicted in Fig. 6-6. After casing is run into the well, an
L-shaped cementing head is fixed to the top of the wellhead to receive the slurry from
the pump. The bottom plug (Fig. 6-7) is first introduced into the well, and cement
slurry is pumped into the well behind the plug. The bottom plug prevents the drilling
mud from mixing with the cement slurry. The plug is then caught just above the bottom
of the wellbore by the float collar, which functions as a one-way valve allowing the
cement slurry to enter the well. Then the pressure on the cement being pumped into the
well is increased until a diaphragm is broken within the bottom plug, permitting the
slurry to flow through it and up the outside of the casing string.
After the proper volume of cement is pumped into the well, a top plug (Fig. 6-7)
is pumped into the casing, pushing the remaining slurry through the bottom plug. Once
the top plug reaches the bottom plug, the pumps are turned off, and the cement is
allowed to set.
Example: You are required to obtain the volume of cement slurry with the
information provided.
Bit Size = 20 in
Casing OD = 16 in
Cement Height = 600 m
Cement Weight = 50 kg/sack
Cement Specific Gravity = 3.14
CaCl2 Concentration = 2% weight of cement
CaCl2 Specific Gravity = 1.83
Chapter 6 Well Completion ·117·

Fig. 6-6 Cementing operation

Fig. 6-7 Top plug and bottom plug


·118· Petroleum Production Technology

Water Cement Ratio = 20 L water per sack of cement


Excess Factor = 1.5
Each sack of cement needs to mix with CaCl2 and water to produce slurry. Let’s
first look at the volume of slurry produced by one sack of cement.
Volume of cement per sack = 50kg/3.14kg/L = 15.92 L
Volume of CaCl2 added to cement = 50kg×2%/1.83kg/L = 1/1.83 = 0.55 L
Volume of water added to cement = 20 L
Therefore,
Volume of slurry produced by one sack of cement = 15.92 L+0.55 L+20 L=36.47L
The cross sectional area of annulus = 0.25×3.14×[(20in)2–(16in)2]=0.073m2
The slurry volume required = 0.073m2×600m×1.5 = 65700L
Number of sacks required = 65700/36.47 = 1801

6.1.5 Perforating a cased well


Since the pay zone is now sealed off by the production casing and cement, perforations
must be made in order for the oil or gas to flow into the wellbore. Perforations are
simply holes that are made through the casing, cement and extend some distance into
the formation, as illustrated in Fig. 6-8.
Bullet guns were used as the first commercial perforating devices. A hardened
steel bullet was fired from a short-barrel gun powered by a gas-producing explosive.
These guns first saw commercial use in the early 1930s. Wall thickness, hardness of the
casing and hardness of the formation limit bullet perforating. Bullet guns are still used
in some applications, usually in soft formations for deep penetration or brittle
formations in which the shattering produced by the bullet can help break down the
formation around the perforation.

Fig. 6-8 Perforation process


Chapter 6 Well Completion ·119·

During the 1930s and 1940s, work in the area of shaped charges progressed in the
military arena. The bazooka, with its armor-piercing charges, was one of the first
large-scale uses of the technology pioneered by Henry Mohaupt and others. This
technology was accepted by the oil industry in early 1950s and became the most
widely-used perforating method by the late 1950s.
The charges are arranged in a tool called a gun that is lowered into the well
opposite the producing zone. Fig. 6-9 presents the components of a jet perforating gun:
the charge carrier, a series of explosive shaped charges, a detonating cord that connects
each charge to a detonator, and an electric circuit to the surface that controls initiation
of perforating sequence. Usually the gun is attached to wire line. When the gun is in
position, the charges are fired by electronic means from the surface. Thus the detonator

Fig. 6-9 Configuration of perforation gun


·120· Petroleum Production Technology

starts a chain reaction that successively initiates the firing of the cord, the primer, and
finally the main explosive. A loaded gun is shown in Fig. 6-10. The tool is retrieved
after the perforations are made. Perforating is usually performed by a service company
that specializes in this technique.
Shaped charges (Fig. 6-11) accomplish penetration by creating a jet of high-
pressure, high-velocity gas. The explosive energy of the detonation is focused in one
direction by the conical liner. This reflects a lot of the energy back into a narrow pulse.
The relatively thin charge liner also plays a critical role by systematically collapsing
and emerging as a high-velocity jet of fluidized metal particles. The pulse moves out at
around 30000 ft/s and generates pressures between 5 and 15 million psi. This pressure
deforms the casing and crushes the cement and formation. The amount of explosive
used is actually small, typically in the range of 6 to 32 grams, although smaller charges
are available for very small-diameter casing and larger charges can be used for big
well-bore sizes.

Fig. 6-10 Loaded perforating gun

Fig. 6-11 Structure of a shaped charge


Chapter 6 Well Completion ·121·

A number of different methods are available for running the gun into the well.
These approaches are schematically illustrated in Fig. 6-12. ① For the first option, the
gun is conveyed by tubing. After perforating the well, the gun is dropped to the bottom
of well. If some misfire occurs, the tubing string has to be pulled. ② The second
option is to convey the gun with drill pipe. In this case, the perforating interval has to
be killed before the gun retrieval, thus resulting in possible formation damage. ③ The

Fig. 6-12 Deployment of perforating gun


·122· Petroleum Production Technology

third option is to convey the gun with wire-line before the tubing is installed. In this
scenario, the well is usually killed before perforating, thus leading to possible
formation damage. ④ The fourth option is to convey the gun with wire-line after the
tubing has been installed. This method limits the gun size because the gun has to pass
through the tubing. The benefit lies in that underbalanced condition can be easily
achieved, thus eliminating the risk of formation damage by fluid in the well.
For each option there are multiple variations. It can be difficult to quantify each of
these advantages and disadvantages when trying to decide what method to use.
Quantifying the relative cost and time is reasonably straightforward. Quantifying the
productivity differences between different sizes of guns can also be made. However,
field engineers often refer to past experiences for gun deployment.

6.2 Upper completion

Production engineers routinely carry out the following tasks related to upper
completion: ① Design of well structure; ② Selection of packer; ③ Tubing movement
calculations; ④ Selection of tubing size; ⑤ Selection and design of artificial lift
methods. Artificial lift will be introduced in chapter 9.

6.2.1 Options for upper completion


The common structures for upper completion are illustrated in Fig. 6-13. ① For
the tubingless completion, no tubing is installed in the well. Fluid flow through casing
to reach surface. ② For the tubing completion without packer, tubing is run into the
well but no packer is installed. Fluid can flow through tubing and tubing/casing
annulus to surface. ③ For the tubing completion with packer, both tubing and packer
are run into the well. The packer is a sealing device that isolates and contains produced
fluids and pressures within the wellbore to protect the casing and other formations
above or below the producing zone. This is essential to the basic functioning of most
wells. ④ For the dual tubing completion, two tubing strings are run into the well. This
design is necessary for producing incompatible fluids from two reservoirs. However,
the small tubing size may restrict flow rates.
Chapter 6 Well Completion ·123·

Fig. 6-13 Options for upper completion

The simplest method consists of producing straight up the casing with no tubing.
This option is cheap and has a large flow area. But fluid flow is difficult to control, and
the casing may be corroded. In most situations, this completion is considered unsafe
due to a lack of barriers in the event of a problem. The second option has some distinct
advantages. The flow may go either up the tubing, up the tubing-casing annulus or up
both. It is therefore very useful in low-pressure gas wells where the flow area can be
switched to overcome liquid loading problems.
For naturally-flowing wells, especially those flowing at moderate or high
pressures, additional barriers to prevent hydrocarbon escape are employed. In the third
option, the tubing is sealed with a packer. Therefore in case the tubing develops a leak,
the casing can withstand the pressure. Thus annulus pressure will be detected, the well
shut-in and the tubing replaced. Replacing tubing is a much easier operation than
replacing casing. Such a completion is very common offshore, where the consequences
of a leak are more severe due to the proximity of staff to the well.
·124· Petroleum Production Technology

The fourth option in Fig. 6-13 employs a dual string. Clearly more complex than
the other options, there are however some useful advantages. This option is used in
low-to-moderate rate wells where there are multiple stacked reservoirs. Flow from the
two intervals is separately produced, controlled and measured. Therefore any problems
with incompatible fluids are avoided. These completions can be very useful if the
reservoir intervals are very different in productivity, pressure or fluids. Rates are
however usually lower than the equivalent single-tubing commingled producer due to
the size limit for two parallel strings inside the casing.

6.2.2 Christmas tree and tubing hanger


A example of well structure is presented in Fig. 6-14. Starting from the top of the well,
a Christmas tree sits on top of the wellhead. The tree is designed to control production
or injection. It is the primary means of shutting in the well. Vertical access through the
tree is possible for logging or other interventions. These operations can be performed
on a flowing well through temporary pressure control equipment installed above the
swab valve. Most wells use a Christmas tree of some form, including subsea wells.

Fig. 6-14 An example of well completion


Chapter 6 Well Completion ·125·

The required pressure rating of the tree is a critical completion decision. It should
be rated above the maximum anticipated pressure for the life of the well. Vertical and
horizontal Christmas trees are available. The difference between the vertical and
horizontal tree is in the position of the valves. In a vertical tree, the master valves are in
the vertical position and inline with the tubing, whilst in a horizontal tree, they are
horizontal and away from the casing (Fig. 6-15). Horizontal trees are preferred for
subsea wells. For subsea operations, the flexibility of being able to install the tree at a
number of different times can be useful as running a subsea tree is weather dependent.

Fig. 6-15 Vertical and horizontal trees

A conventional land or platform tree often consists of two master valves, a wing
valve and a swab valve. A second wing valve can be useful for pumping operations
such as stimulation or chemical treatments. For many platform wells, one of the master
valves and the wing valve are hydraulically actuated and connected to the platform
shut-down system. The swab valve is almost always manual. This provides the
sensitivity to count turns when closing the valve after a through-tubing intervention.
The tubing hanger is a metal tool that supports the tubing. It is either installed
inside the wellhead, or it sits inside the tree for certain types of tree. The tubing hanger
·126· Petroleum Production Technology

connects to the tree via seals, and to the tubing below via a screwed thread (Fig. 6-16).
The tubing hanger usually has penetrations for control lines, down-hole gauge lines and
chemical injection lines. For instance, four control lines protrude through the tubing
hanger in Fig. 6-17.

Fig. 6-16 Cutaway of well head and tubing hanger

Fig. 6-17 Tubing hanger


Chapter 6 Well Completion ·127·

Below the hanger comes the tubing. The tubing has to be designed to withstand
high pressure, and sometimes high temperature. The production fluids are often
corrosive and the tubing is usually made out of corrosion-resistant alloys such as
stainless steel, especially for critical and high-rate wells. The tubing, like casing, comes
in joints typically 30-40 ft long and is screwed together on the rig.

6.2.3 Safety valve


A vital safety valve named subsurface safety valve (SSSV) is always installed on
tubing often around 100 ft from surface. It prevents the uncontrolled release of reservoir
fluids in the event of a possible surface disaster. These valves are commonly flapper
valves which open downwards such that the flow of wellbore fluids tries to push it shut,
while pressure from the surface pushes it open. When closed, it will isolate the
reservoir fluids from the surface.
Most down-hole safety valves are controlled hydraulically from the surface,
meaning they are opened using a hydraulic connection linked directly to a well control
panel. The opening and closing mechanisms are depicted in Fig. 6-18. When hydraulic
pressure is applied down a control line, the hydraulic pressure forces a sleeve within
the valve to slide downwards. This movement compresses a large spring and pushes the

Fig. 6-18 Subsurface safety valve


·128· Petroleum Production Technology

flapper downwards to open the valve. When hydraulic pressure is removed, the spring
pushes the sleeve back up and causes the flapper to shut. In this way, it is failsafe and
will isolate the wellbore in the event of loss of the wellhead.
Below the safety valve comes a range of optional equipment. Shown in Fig. 6-14
are three types of mandrel. One for the injection of gas, one for a pressure gauge and
the other for chemical injection. The types of chemicals injected could include scale
inhibitors, methanol for hydrate inhibition and corrosion inhibitors. Another optional
piece of equipment is a sliding side door for the circulation of fluids into and out of the
tubing.
In the example (Fig. 6-14), the tubing is anchored and sealed to the casing with a
production packer. Sometimes above the packer, an expansion device such as the
polished bore receptacle is used to allow for thermal expansion or contraction of the
tubing. Below the packer, there is usually a tailpipe with a nipple profile for the setting
of plugs, temporary gauges or down-hole chokes. Nipple profiles are also found inside
the tubing hanger, immediately above a safety valve. The nipple (Fig. 6-19) is a
permanent part of the completion.

Fig. 6-19 Two common no-go nipple profile


Chapter 6 Well Completion ·129·

6.2.4 Production packer


In addition to providing a seal between the tubing and casing, other functions of a
packer are as follows: ① to prevent down-hole movement of the tubing string; ② to
support some of the weight of the tubing; ③ to protect the annular casing from
corrosion from produced fluids and high
pressures; ④ to provide a means of
separation of multiple producing zones;
⑤ to hold well-servicing fluid (kill fluids
and packer fluids) in the casing annulus.
Production packers can be set by
inflation or swelling. Inflatable packers are
set hydraulically or mechanically with tension
or compression. The key components of an
inflatable packer include slip, cone, packing
element and body or mandrel (Fig. 6-20).
The slip is a wedge-shaped device with
teeth on its face, which penetrate and grip Fig. 6-20 Components on an inflatable packer
the casing wall when the packer is set. The
cone is beveled to match the back of the slip and forms a ramp that drives the slip
outward and into the casing wall when setting force is applied to the packer. Once the
slips have anchored into the casing wall, additional applied setting force activates the
packing-element system and creates a seal between the packer body and the inside
diameter of the casing.
Inflatable packers can be further classified into retrievable type and permanent
type. Once set, permanent packers can not be retrieved from the well. A milling tool is
required to remove a permanent packer from a well. The tool destroys the packer’s
slips, then the packer can be retrieved to surface. The retrievable packer may or may
not be reset, but removal from the wellbore normally does not require milling.
Retrieval is usually accomplished by some form of tubing manipulation. This may
require rotation or pulling tension on the tubing string.
The permanent packer is fairly simple and generally offers higher performance in
both temperature and pressure ratings than the retrievable packer. The permanent
packer has very few components, therefore being less costly than other packers. In
·130· Petroleum Production Technology

most instances, it has a smaller outside diameter, offering greater running clearance
inside the casing string than retrievable packers. The smaller outer diameter and the
compact design of the permanent packer help the tool negotiate through tight spots and
deviations in the wellbore. The permanent packer also offers the largest inside diameter
to be compatible with larger-diameter tubing. Due to these factors, permanent packers
have become the most widely-used packers in the world.
Permanent packers can be set quickly and accurately by wire-line at a certain
depth. After the packer is set, a production seal assembly and production tubing are run
into the well. Once they are run into the well, the tubing seals engage the packer’s
internal seals, tubing length is then adjusted at surface, and the completion is finished.
Swell packers are recent development in packer technology. They provide very
effective seal in both open and cased wells. They swell based on two types of
mechanisms: Water swelling and oil swelling. In the case of water swelling elastomers,
swelling process is based on the principle of osmosis. Water enters the rubber matrix
and swells the element until the equilibrium is achieved. Elastomer and the salinity of
surrounding fluid/water are very important factors for osmosis process. Changes in
down-hole conditions and fluid properties can reverse the swelling process.
Oil swelling elastomers swell by the diffusion process. Rubber molecules absorb
the hydrocarbon molecule, thus causing elastomers to stretch. Cross-linked polymer
network of swelling packer rubber traps the hydrocarbon molecules due to their natural
affinity. Reversible process is not possible. Unlike the inflatable packer, swelling
packers deployment time can take few hours to several weeks. Swelling packers gained
market share for simple handling and proven efficiency.

6.2.5 Smart completion


A more advanced completion is the well equipped with remote down-hole flow
controls. These are often called smart wells or intelligent wells. The configuration
shown in Fig. 6-21 is a smart completion for a cased and perforated well. The valves
are either electrically or hydraulically controlled from surface via the control cables.
Packers isolate each section of the reservoir or different reservoirs. The valves can be
the on-off type or for added control and complexity, can be variable.
Chapter 6 Well Completion ·131·

Fig. 6-21 An example of smart completion

Most smart wells will also deploy gauges for pressure and temperature
measurements both inside and outside of the tubing at each control valve. The gauge
signals are usually multiplexed into the valve control signals. Smart completions are
very expensive, therefore they are only chosen for the most important wells in the field.
Some subsea wells also employ smart completion because well intervention is difficult
and costly.
Chapter 7 Well Inflow Performance
In field practice, it is crucial to monitor and diagnose day-to-day well production
performances through the life cycle of the field. The well production performance is
quantified with well productivity index and well inflow performance relation (IPR).

7.1 Well productivity index


Productivity index (PI or J) is the measurement of a well’s ability to produce liquid. PI
is the ratio of the total liquid flow rate to pressure drawdown. It is expressed as:
Q
PI = (7-1)
Pr − Pwf

where,
PI = productivity index, stb/(d ⋅ psi);
Q = liquid flow rate, stb/d;
Pr = reservoir pressure, psi;
Pwf = bottom-hole flowing pressure, psi.
The PI concept is widely used because it is very convenient to evaluate well
performance with productivity index. PI is also very useful for comparison of wells’
performance in the same field. The productivity index is generally measured with a
production test on the well. The well is shut in until the static reservoir pressure is
reached. The well is then allowed to produce at a constant flow rate and stabilized
bottom-hole pressure. The bottom-hole pressure should be recorded continuously. The
productivity index can be calculated with the well test data.
It is important to note that the productivity index is a valid measurement of the
well productivity only if the well is flowing under pseudo-steady state. Therefore, it is
essential that the well is allowed to flow at a constant flow rate for a sufficiently long
time to reach the pseudo-steady state. Rearrange the definition of PI and we can obtain

Q = PI ( Pr − Pwf ) (7-2)

Assume PI is constant and we can plot a graph for oil flow rates at varied well
Chapter 7 Well Inflow Performance ·133·

flowing pressure, as shown in Fig. 7-1. The relationship between flow rates and well
flowing pressure is referred to as inflow performance relationship (IPR). The
maximum flow rate on the plot is named absolute open flow potential (AOF). AOF is
only a conceptual parameter, because it is impossible to produce the well at zero well
flowing pressure.

Fig. 7-1 Straight-line IPR

The straight-line IPR has certain limitations. When the well flowing pressure is
below the bubble point pressure of reservoir oil, additional drawdown may not yield a
proportional increase in liquid production rate. This is due to the presence of free gas in
the reservoir. Therefore, the straight-line IPR over-predicts production rate while well
flowing pressure is lower than bubble point pressure. On the other hand, for a
multi-layered reservoir, the inflow relationship is very complex and it does not follow a
straight line. The different layers may be under different pressures. Some layers may
not flow when the drawdown is small. Under high drawdown, additional layers begin
to flow into the well. As a result, the productivity of the well changes when the
drawdown is changed. Several empirical methods have been developed to generate
well IPR. They are more realistic than the straight line IPR, because these methods take
into account the gas effect at low pressure.

7.2 Vogel IPR

In 1968, Vogel proposed an IPR for saturated reservoirs through a computer and
database study on a wide range of parameters for solution-gas-drive reservoirs. Vogel
·134· Petroleum Production Technology

IPR requires only one well test on static reservoir pressure, plus a single flow rate with
corresponding well flowing pressure. The Vogel IPR was originally developed for
reservoirs under solution gas drive, but it can be used for any type of reservoir in
practice.
2
Q P ⎛P ⎞
= 1 − 0.2 wf − 0.8 ⎜ wf ⎟ (7-3)
AOF Pr ⎝ Pr ⎠

where,
Pr = reservoir pressure, psig;
Pwf = bottom-hole flowing pressure, psig①;
AOF = absolute open flow potential, stb/d.
Example: A well is producing from a saturated reservoir. The reservoir pressure
and bubble point pressure are both at 3000 psig. A stabilized production test was
conducted at well flowing pressure of 2000 psig, and the production was 500 stb/d. ①
calculate well productivity index; ② calculate the well flow rate at a well flowing
pressure of 1500 psig, assuming constant PI; ③ calculate the well flow rate at a well
flowing pressure of 1500 psig with Vogel method; ④ construct Vogel IPR.

Q 500
PI = = = 0.5(stb d ⋅ psi)
Pr − Pwf 3000 − 2000


Q = PI ( Pr − Pwf ) = 0.5 ( 3000 − 1500 ) = 750(stb / d)

③ First, we need to rearrange Vogel IPR to calculate AOF based on the


production test data.
Q
AOF = 2
P ⎛P ⎞
1 − 0.2 wf − 0.8 ⎜ wf ⎟
Pr ⎝ Pr ⎠
500
= 2
= 978(stb / d)
2000 ⎛ 2000 ⎞
1 − 0.2 × − 0.8 × ⎜ ⎟
3000 ⎝ 3000 ⎠

Next, we calculate the flow rate at Pwf = 1500 psig:


————————
① psig refers to gauge pressure, 1psig=6.895kPa+atmospheric pressure.
Chapter 7 Well Inflow Performance ·135·

⎡ Pwf ⎛ Pwf ⎞ ⎤
2
Q = AOF ⎢1 − 0.2 ⋅ − 0.8 ⋅ ⎜ ⎟ ⎥
⎢ Pr ⎝ Pr ⎠ ⎥
⎣ ⎦
⎡ 1500 ⎛ 1500 ⎞ ⎤
2
= 978 ⎢1 − 0.2 × − 0.8 × ⎜ ⎟ ⎥ = 685(stb / d)
⎣⎢ 3000 ⎝ 3000 ⎠ ⎦⎥

It can be seen the straight line IPR over-predicts production rate.


④ Now we assume a series of well flowing pressures, and calculate the
corresponding flow rates with Vogel method, as shown in Tab. 7-1. The IPR is plotted
with the calculation results (Fig. 7-2).

Tab. 7-1 Calculated results for Vogel IPR


Pwf /psig Liquid rate/(stb/d)

3000 0

2500 272

2000 500

1500 685

1000 826

500 924

0 978

Fig. 7-2 Vogel IPR for a saturated reservoir

Vogel method has to be modified for under-saturated reservoirs. For under-


saturated reservoirs, their IPR curves are formed by two parts. Above bubble point
pressure, IPR is a straight line because only single-phase liquid flows into the well.
·136· Petroleum Production Technology

Below bubble point pressure, the IPR curve is actually a Vogel IPR that starts at bubble
point pressure. This analysis is illustrated in Fig. 7-3. There are two possible scenarios,
according to the pressure measurement of production tests.

Fig. 7-3 Vogel IPR for under-saturated reservoir

Case 1: The production tests is conducted at a well flowing pressure above bubble
point pressure. For this case, we first need to obtain PI and the flow rate at bubble point
pressure. Then IPR can be constructed with the modified Vogel method. These steps
are as follows.
Step 1: Obtain PI with production test data.

Q
PI =
Pr − Pwf

Step 2: Obtain flow rate at bubble point pressure.

Qb = PI ( Pr − Pb )

Step 3: Construct IPR for well flowing pressure above bubble point pressure.

Q = PI ( Pr − Pwf )

Step 4: Construct IPR below bubble point pressure with modified IPR

PIPb ⎡ ⎛ Pwf ⎞ ⎤
2
Pwf
Q = Qb + ⎢1 − 0.2 − 0.8 ⎜ ⎟ ⎥
1.8 ⎢ Pb ⎝ Pb ⎠ ⎥
⎣ ⎦
Chapter 7 Well Inflow Performance ·137·

Example: Construct Vogel IPR with the given data. Reservoir pressure = 3000
psig ; Bubble point pressure = 1600 psig ; Liquid rate = 500 stb/d when well flowing
pressure = 2000 psig.
Following the steps described earlier, we can obtain the results below:
PI = 0.5 stb/(d ⋅ psi)
Qb = 700 stb/d
We assume various well flowing pressures above bubble point pressure, and the
liquid rate can be calculated with the straight-line IPR. The results are summarized in
Tab. 7-2.
Tab. 7-2 Vogel IPR above bubble point pressure

Pwf /psig Liquid rate/(stb/d)

1600 700

2000 500

2500 250

3000 0

We assume various well flowing pressures below bubble point pressure, and
calculate corresponding liquid rates with the modified Vogel IPR.

PIPb ⎡ ⎛ Pwf ⎞ ⎤
2
Pwf
Q = Qb + ⎢1 − 0.2 − 0.8 ⎜ ⎟ ⎥
1.8 ⎢ Pb ⎝ Pb ⎠ ⎥
⎣ ⎦

The results are summarized in tab. 7-3, and the complete IPR is plotted in Fig. 7-4.
Tab. 7-3 Vogel IPR below bubble point pressure

Pwf /psig Liquid rate /(stb/d)

0 1144

300 1115

600 1061

900 982

1200 878

1600 700
·138· Petroleum Production Technology

Fig. 7-4 Vogel IPR for under-saturated reservoir

Case 1: production test pressure above bubble point pressure


Case 2: The production test is conducted at a well flowing pressure below bubble
point pressure. For this case, we can follow the steps below to construct IPR.
Step 1: Obtain PI with production tests data.
Q
PI =
Pb ⎡ ⎛P ⎞ ⎤
2
P
( Pr − Pb ) + ⎢1 − 0.2 wf − 0.8 ⎜ wf ⎟ ⎥
1.8 ⎢ Pb ⎝ Pb ⎠ ⎥⎦

Step 2: Obtain flow rate at bubble point pressure.


Qb = PI ( Pr − Pb )

Step 3: Construct IPR for well flowing pressure above bubble point pressure.
Q = PI ( Pr − Pwf )

Step 4: Construct IPR for well flowing pressure below bubble point pressure.

PIPb ⎡ P ⎛P ⎞ ⎤
2
Q = Qb + ⎢1 − 0.2 wf − 0.8 ⎜ wf ⎟ ⎥
1.8 ⎢ Pb ⎝ Pb ⎠ ⎥⎦

Example: Construct Vogel IPR with the given data. Reservoir pressure = 3000
psig ; Bubble point pressure = 2400 psig ; Liquid rate = 500 stb/d when well flowing
pressure = 2000 psig.
Following the steps described previously, we can obtain the results below.
PI = 0.515 stb/d ⋅ psi
Chapter 7 Well Inflow Performance ·139·

Qb = 309 stb/d
For well flowing pressures above bubble point pressure, IPR follows the straight line
relation. For well flowing pressures below bubble point pressure, we assume well
flowing pressures and calculate corresponding flow rates with the equation:

PIPb ⎡ P ⎛P ⎞ ⎤
2
Q = Qb + ⎢1 − 0.2 wf − 0.8 ⎜ wf ⎟ ⎥
1.8 ⎢ Pb ⎝ Pb ⎠ ⎥⎦

The results are given in Tab. 7-4, and IPR is plotted in Fig. 7-5 accordingly.
Tab. 7-4 Calculated results for Vogel IPR
Pwf /psig Liquid rate /(stb/d)
0 996
300 970
600 927
900 867
1200 790
1500 695
1800 584
2000 500
2400 309
2700 155
3000 0

Fig. 7-5 Vogel IPR for an under-saturated reservoir

case 2: production test pressure below bubble point pressure


·140· Petroleum Production Technology

7.3 Fetkovich IPR

Fetkovich proposed the following method to predict gas well IPR. This equation is also
valid for oil wells.
2 n
Q = c( Pr2 − Pwf ) (7-4)

There are two empirical parameters in the Eq (7-4) (n and c). Obviously, we
need to have at least two production tests to solve for the two parameters. More tests
should be done to guarantee data quality. It is easier to solve for the two parameters by
rewriting the IPR as

logQ = logc + n lg( Pr2 − Pwf


2
) (7-5)

Example: Two production tests were conducted on a well, and the results are
given below. Construct Fetkovich IPR with the data provided. Reservoir pressure =
3000 psig ; Test 1: Flow rate = 800 stb/d when well flow pressure = 1000 psig ; Test 2:
Flow rate = 500 stb/d when well flow pressure = 2000 psig.
With the production test data, the following relations are established:

lg800 = lgc + n lg(30002 − 10002 )

lg 500 = lg c + n lg(30002 − 20002 )

We can find n = 1, and c = 0.0001. Assume a series of well flowing pressures, and the
corresponding flow rates can be calculated (Tab. 7-5).

Tab. 7-5 Caculated results for Fetkovich IPR Continued


Well flowing pressure/psig Flow rate/(stb/d)

0 900

300 891

600 864

1000 800

1300 731

1600 644

2000 500
Chapter 7 Well Inflow Performance ·141·

Continued
Well flowing pressure/psig Flow rate/(stb/d)

2300 371

2600 224

3000 0

The Fetkovich IPR is plotted in Fig. 7-6.

Fig 7-6 An example of Fetkovich IPR

7.4 Productivity of Horizontal Wells

Horizontal wells were first drilled in the 1950s, but gained great popularity from the
1980s onwards, as directional drilling technology progressed and the cost reduced.
Horizontal wells have several potential advantages over vertical or deviated wells.
Horizontal wells usually yield higher production rates than vertical wells. The long
horizontal section allows increased exposure to the reservoir. Nowadays, horizontal
sections many kilometers in length are routine in many fields. Because the PI is a
function of the length of reservoir drained by a well, horizontal wells can give higher
productivities in laterally extensive reservoirs.
Horizontal wells have a large potential to connect laterally discontinuous features
in heterogeneous or discontinuous reservoirs. If the reservoir is faulted or fractured, a
horizontal well may connect a series of fault blocks or natural fractures in a manner
which would require many vertical wells. The ultimate recovery of a horizontal well is
likely to be significantly greater than a single vertical well (Fig. 7-7).
·142· Petroleum Production Technology

Fig. 7-7 A horizontal well in fractured reservoir

The third benefit of horizontal wells is to reduce the effects of coning. For
example, a horizontal producing well may be placed far away from the advancing
oil-water contact during water drive. An additional advantage is that if the PI for the
horizontal well is larger, then the same oil production can be achieved at much lower
drawdown, therefore minimizing the effect of coning. As a result, oil production is
achieved with significantly less water production, which reduces processing costs and
assists in maintaining reservoir pressure. Horizontal wells have a particularly strong
advantage in thin oil columns (less than 40m thick), which would be prone to coning
if developed with vertical wells. The fluid interface near the horizontal well is often
referred to as cresting rather than coning, due to the shape of the interface (Fig. 7-8).

Fig. 7-8 Gas cresting induced by a horizontal well

As an estimate of the potential benefit of horizontal well, the productivity


improvement factor (PIF) compares the initial productivity of a horizontal well to that of
a vertical well in the same reservoir, during early radial flow. The factor is defined as:

L kv
PIF = (7-6)
h kh

where,
Chapter 7 Well Inflow Performance ·143·

L = length of the reservoir;


h = thickness of the reservoir;
kh = horizontal permeability of the reservoir;
kv = vertical permeability of the reservoir.
The reservoir geometry and reservoir property have very important influences on the
benefit of horizontal wells over a vertical well, as demonstrated by the following examples.
Each of these examples assumes that the reservoir is a block with homogeneous properties.
For the case1 shown in Fig. 7-9, the reservoir length equals thickness, and the vertical
permeability equals horizontal permeability. In this case, a horizontal well does not yield
higher production than a vertical well.

Fig. 7-9 Evaluation of PIF(case 1)


kv=kh; PIF=1.0

For the case2 in Fig. 7-10, the reservoir vertical permeability equals horizontal
permeability, but the reservoir length is significantly greater than its thickness. A
horizontal well yields much higher production than a vertical well.

Fig. 7-10 Evaluation of PIF(case 2)


kv=kh; PIF=4.0
·144· Petroleum Production Technology

In the case of very low vertical permeability(see Fig. 7-11), the horizontal well
actually produces at a lower rate than the vertical well. For these scenarios, the ultimate
recovery achieved with the horizontal well is unlikely to be different to that with the
vertical well. The major benefit is the accelerated production achieved by horizontal wells.

Fig. 7-11 Evaluation of PIF (case 3)


kv=0.01 kh; PIF=0.4

Actually in high permeability reservoirs, there is a diminishing return of


production rate on the length of well drilled, due to increasing friction pressure drop
with increasing well length. This effect is shown schematically in Fig. 7-12.

Fig. 7-12 Production rate versus length of horizontal well

In 1984, Borisov proposed the following equation to estimate productivity of a


horizontal well :

0.00708hkh
Jh = (7-7)
⎡ 4r h h ⎤
μo Bo ⎢ln eh + ln ⎥
⎣ L L 2πrw ⎦

where,
Chapter 7 Well Inflow Performance ·145·

Jh= productivity index of a horizontal well;


h = reservoir thickness, ft;
kh = reservoir horizontal permeability, mD;
L = length of horizontal well, ft;
rw = wellbore radius, ft;
reh = drainage radius of horizontal well, ft.
In 1991, Joshi proposed the equation below to estimate the productivity of a
horizontal well in isotropic reservoir :
0.00708hkh
Jh = (7-8)
⎡ h h ⎤
μo Bo ⎢ln R + ln ⎥
⎣ L 2rw ⎦

where,
2
⎛L⎞
a + a2 − ⎜ ⎟
⎝2⎠
R=
⎛L⎞
⎜ ⎟
⎝2⎠

and,
0.5
⎡ 4⎤
⎛L⎞ ⎛ 2r ⎞
a = ⎜ ⎟ ⎢0.5 + 0.25 + ⎜ eh ⎟ ⎥
⎝ 2 ⎠⎢ ⎝ L ⎠ ⎥
⎣ ⎦

Joshi also proposed a method to incorporate reservoir anisotropy :


0.00708hkh
Jh = (7-9)
⎡ B2h h ⎤
μo Bo ⎢ln R + ln ⎥
⎣ L 2rw ⎦

where,

kh
B= (7-10)
kv

Example: A 5000ft horizontal well is completed in a 100 acre plot. The


reservoir is anisotropic with the following data available. Calculate well productivity
·146· Petroleum Production Technology

with Joshi method. kv = 10 mD ; kh = 100 mD ; h = 100 ft ; rw = 6 in = 0.5 ft ; Bo =


1.1 bbl/stb ; μo = 10 cP .
First, we need to calculate the drainage radius, assuming the well drains a circular
field :

43560 × 100
reh = = 1178 (ft)
π

and,

a = 2558 ft; R = 1.24; B = 3.162

The PI of the horizontal well:


0.00708hkh
Jh =
⎡ B2 h ⎛ h ⎞⎤
μo Bo ⎢lnR + ln ⎜ ⎟⎥
⎣ L ⎝ 2rw ⎠ ⎦
0.00708 × 100 × 100
= = 5.665[stb (d ⋅ psi)]
⎡ 10 × 100 100 ⎤
10 × 1.1 × ⎢ln1.24 + × ln
⎣ 5000 2 × 0.5 ⎥⎦
Chapter 8 Fluid Flow in Wells and Nodal Analysis
The produced fluid from a producing well often goes through a complex well trajectory
to reach surface, then through surface pipes to arrive at treatment and processing
facilities. When fluid flows in wells and pipes, it has to overcome three types of
pressure losses: hydrostatic pressure loss, friction pressure loss, and acceleration head.
Hydrostatic pressure loss weighs around 80% total pressure loss in oil wells, while
friction loss may become dominant in gas wells. For production engineers, the
prediction of pressure loss in wells lays the foundation for production optimization and
artificial lift design.
Production wells usually produce gas, oil and water simultaneously. The
commingled flow of multiple phases is referred to as multiphase flow. Single-phase flow
and some basic concepts will be introduced first, before moving on to multiphase flow.

8.1 Single-phase flow in pipes

While fluid flows through tubing or pipe, it has to overcome elevation pressure loss,
friction pressure loss and kinetic pressure loss. These pressure losses are defined as
follows:
ΔPe = ρgΔh (8-1)

2 fρv 2 L
ΔPf = (8-2)
D
ρv 2
ΔPa = (8-3)
2
where,
ΔPe = elevation (or hydrostatic) pressure loss, Pa;
ΔPf = friction pressure loss, Pa;
ΔPa = acceleration (or kinetic) pressure loss, Pa;
Δh = elevation change, m;
f = Fanning friction factor;
L = pipe length, m;
·148· Petroleum Production Technology

ρ = fluid density, kg/m3;


v = fluid velocity, m/s;
D = pipe inside diameter, m.
Reynolds number represents the ratio of inertial force to viscous force. Reynolds
number is defined as:
ρvD
Re = (8-4)
μ

where,
Re = Reynolds number;
ρ = fluid density, kg/m3;
μ = fluid viscosity, kg/(m ⋅ s);
v = fluid velocity, m/s;
D = pipe inside diameter, m.
The Reynolds number indicates the level of turbulence in fluid. Laminar flow
occurs when Reynolds number is below 2000 (some may use 2200 or 3000 as the
criterion). When Reynolds number exceeds 2000, the flow transforms from laminar to
turbulent flow.
The friction factor must be determined in order to calculate friction pressure loss.
In practice, two friction factors are frequently used: Fanning friction factor (denoted as
f) and Moody friction factor (denoted as fm). The relation between Moody friction
factor and Fanning friction factor is expressed as:
fm = 4 f (8-5)

Under laminar flow, the Fanning friction factor is related to Reynolds number only:
16
f = (8-6)
Re
For turbulent flow, the Colebrook-White equation is widely used to calculate Moody
friction factor, also know as Darcy friction factor. The equation is implicit, thus an
iteration process is required to obtain the friction factor. The equation is expressed as:

1 ⎛ 18.7 ⎞
= 1.74 − 2lg ⎜ 2δ + ⎟ (8-7)
fm ⎜ Re f m ⎟⎠

and,

δ=ε/D
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·149·

where,
ε = absolute pipe roughness;
δ = relative pipe roughness;
The Fanning friction factor can be estimated with Chen correlation. This
correlation takes an explicit form and the result is similar to that with Colebrook-White
equation. Chen equation is presented here:

1 ⎧⎪ δ 5.0452 ⎡ δ1.11 ⎛ 7.149 ⎞


0.898 ⎤ ⎫

= −4lg ⎨ − lg ⎢ +⎜ ⎟ ⎥⎬ (8-8)
f 3.7065 Re ⎢⎣ 2.8257 ⎝ Re ⎠ ⎥⎦ ⎭⎪
⎩⎪

According to Chen equation and Colebrook-White equation, friction factor is


influenced by Reynolds number and pipe roughness. Moody friction factors are plotted
in Fig. 8-1. Three conclusions can be drawn based on the chart: ① Under laminar
flow, friction factor is a function of Reynolds number. ② Under turbulent flow, as
Reynolds number increases, the friction factor first decreases then levels off. ③
Under turbulent flow, high pipe roughness yields a high friction factor.

Fig. 8-1 Moody friction factor chart


·150· Petroleum Production Technology

Example: Water flows in a pipe. With the data provided, calculate friction
pressure loss. Water flow rate = 0.07854 m3/s ; water density = 1000 kg/m3 ; water
viscosity = 0.01 poise ; pipe inner ID = 10 cm ; pipe roughness = 0.1 mm ; pipe length
= 100 m.
Pipe cross sectional area is:

Ap = 0.25πD2 = 0.007854 (m2)

Water velocity is:

v = 0.07854/0.007854 = 10 (m/s)

Reynolds number is:

ρvD 1000 × 10 × 0.1


Re = = = 106
μ 0.001

Flow is clearly turbulent.

Relative roughness = 0.1mm/10cm = 0.001

Moody friction factor = 0.02 (Fig. 8-1)

Fanning friction factor = 0.005 (Eq. (8-8))

Friction pressure loss is:

2 fρv 2 L 2 × 0.005 × 1000 × 102 × 100


ΔPf = = = 106 (Pa)
D 0.1

8.2 Multiphase flow concept

Multiphase flow takes place when more than one phases are present in the flow. Fluid
from a production well is usually a mixture of oil, gas and water. Sometimes solids are
also carried with the fluid, such as sand, wax, and hydrates. Oil and water have
relatively similar properties, therefore they are often regarded as one phase in
multiphase flow concept. Multiphase flow models usually ignore the solids in
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·151·

production stream, because the quantity of solids is usually very low. As such, the four
phases can be simplified into two phases: Liquid and gas.
Liquid-gas flow is much more complicated than the flow of a single phase.
Various flow patterns (or flow regimes) are formed in two-phase flow. Fluids are
distributed differently under different flow patterns, thus greatly affecting the
corresponding pressure loss in wells and pipes. Four flow regimes have been identified
in two-phase flow in a vertical well or pipe: Bubble flow, slug flow, churn flow, and
annular flow, as shown in Fig. 8-2.

Fig. 8-2 Patterns for two-phase flow in a vertical pipe

These flow patterns occur in sequence as the gas flow rate increases for a given
liquid flow rate. When the gas phase first liberates from oil, gas appears as small
bubbles in the liquid phase. This flow pattern is thus named bubble flow. For bubble
flow, gas phase is the dispersed phase and liquid is the continuous phase. For slug flow,
small gas bubbles coalesce into large bubbles that eventually fill the entire pipe
cross-section. Between the large bubbles are slugs of liquid that contain smaller
bubbles of entrained gas. For churn flow, the larger gas bubbles become unstable and
collapse, resulting in a highly turbulent flow pattern with both phases dispersed. For
annular flow, gas flow rate is very high and gas breaks through the center of the pipe,
with liquid coating the surface of the pipe and droplets entrained in the gas phase. Both
phases are continuous in annular flow.
For two-phase flow in a horizontal well or pipe, five flow patterns are recognized as
bubble flow, wavy flow, slug flow, annular flow and mist flow, as presented in Fig. 8-3.
·152· Petroleum Production Technology

Among these patterns, bubble flow and wavy flow occur at relatively low gas flow rate,
while slug flow occurs at medium gas flow rate. When gas flow rate is extremely high,
gas phase occupies the middle of the pipe (namely annular flow), or breaks liquid
phase into small droplets (namely mist flow).

Fig. 8-3 Patterns for two-phase flow in a horizontal pipe


Chapter 8 Fluid Flow in Wells and Nodal Analysis ·153·

For upward flow, gas travels faster than liquid phase due to the difference in
density. This phenomenon is named slippage. The slippage velocity is defined as:

vs = vG − vL (8-9)

where,
vs = slippage velocity;
vG = velocity of gas phase;
vL = velocity of liquid phase.
As gas and liquid travel upwards, slippage becomes more severe, and the volume
fraction of liquid increases in a segment of pipe. This phenomenon is similar to liquid
being held up, while gas travels upwards. The liquid holdup is defined as:
VL AL
HL = = (8-10)
V A

where,
HL = liquid holdup (sometimes denoted as yL);
VL = volume of liquid in a pipe segment;
V = volume of the pipe segment;
AL = pipe cross-sectional area occupied by liquid;
A = pipe cross-sectional area.
Flow pattern is the main factor that influence liquid holdup. Its value can be
measured by experiments or predicted by correlations. On top of liquid holdup, we can
obtain some important parameters in multiphase flow, including mixture density,
superficial phase velocity, and real phase velocity.
Mixture density is defined as:
ρs = ρL H L + ρG (1 − H L ) (8-11)

where,
ρL = density of liquid; ρG = density of gas.
Mixture viscosity is defined as:

μs = μLH L μG(1− H L ) (8-12)

or,

μs = μL H L + μG (1 − H L ) (8-13)
·154· Petroleum Production Technology

where,
μL = viscosity of liquid;
μG = viscosity of gas.
Based on the definitions, we can see two-phase mixture properties are affected by
liquid holdup and slippage. It is obvious that the density of gas-liquid mixture is less
than that of the single phase liquid, thus reducing hydrostatic pressure loss in upward
flow. In practice, the liquid phase is often a mixture of oil and water. Therefore, the
oil-water mixture density and viscosity should be obtained first.
In multiphase models, superficial velocities are often used as an indicator of flow
pattern. They are defined as:
qL
vSL = (8-14)
A

and,
qG
vSG = (8-15)
A

where,
vSL = superficial liquid velocity;
vSG = superficial gas velocity;
qL = liquid flow rate;
qG = gas flow rate.
It can be seen the superficial velocities take into account the whole pipe cross-
sectional area, even though each phase occupies only a fraction of the pipe. The real
phase velocities are:
qL
vL = (8-16)
AH L

and,
qG
vG = (8-17)
A(1 − H L)

where,
vL = real liquid velocity;
vG = real gas velocity.
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·155·

The mixture velocity is expressed as:


qL + qG
vm = = vSL + vSG (8-18)
A

Ansari et al. proposed superficial velocities as indicators for flow patterns. They
plotted a flow-pattern map based on experimental data (Fig. 8-4). It is clear that the
flow patterns are influenced by liquid and gas superficial velocities. When gas velocity
is low, the flow pattern is bubble flow. When gas velocity increases, flow pattern
moves into slug flow and churn flow. At very high gas velocity, gas breaks through
from center of pipe and annular flow is observed.

Fig. 8-4 Superficial velocity and flow pattern

8.3 Multiphase flow models

Numerous models have been developed for prediction of pressure loss of multiphase
flow in wells and pipes. Multiphase flow models fall into two categories: Homogeneous
flow models and separated flow models. Homogeneous models treat multiphase as a
homogeneous mixture and do not consider the effects of liquid holdup or slippage.
·156· Petroleum Production Technology

Therefore, these models are less accurate and are usually calibrated with local operating
conditions in field applications.
Separated-flow models are more realistic than the homogeneous-flow models.
They are usually given in the form of empirical correlations. The effects of liquid
holdup, slippage and flow regimes are considered. The major disadvantage of the
separated flow models is that it is difficult to code them in computer programs, because
most correlations are presented in graphic form. Some of the widely-used models are
summarized in Tab. 8-1.
Tab. 8-1 List of commonly-used multiphase flow models
Type of model Founder of model Year of publication
Homogeneous Model Poettman and Carpenter 1952
Homogeneous Model Fancher and Brown 1963
Homogeneous Model Hagedorn and Brown 1965
Separated Flow Model Duns and Ros 1963
Separated Flow Model Orkiszewski 1967
Separated Flow Model Beggs and Brill 1973

These models are compared in Fig. 8-5. It can be seen the Beggs-Brill method
predicts the highest pressure loss in vertical flow, The Orkiszewski model and the
Hagedorn-Brown model predict the lowest pressure loss. In practice, we run these
models and calibrate the model with production test data for the well being analyzed.

Fig. 8-5 Comparison of multiphase flow models


Chapter 8 Fluid Flow in Wells and Nodal Analysis ·157·

Based on comprehensive comparisons of these models, Ansari et al., Hasan and


Kabir recommended the Hagedorn-Brown method with modifications for near-vertical
flow. The modified Hagedorn-Brown method is an empirical correlation developed on
the basis of the original work of Hagedorn and Brown. The modifications include using
the no-slip liquid holdup when the original correlation predicts a liquid holdup value
less than the no-slip holdup and using the Griffith correlation for the bubble flow
pattern. The original Hagedorn-Brown correlation takes the following steps:
Step 1: Calculate the following parameters.
Liquid velocity number:
4
ρL
N VL = 1.938vSL (8-19)
σ

Gas velocity number:


4
ρL
N VG = 1.938vSG (8-20)
σ

Pipe Diameter number:

ρL
N D = 120.872 D (8-21)
σ

Liquid viscosity number:


4
1
N L = 0.157 μL 3
(8-22)
σ ρL

where,
σ = liquid-gas interfacial tension;
D = pipe inner diameter.
Liquid contact number:

CN L = 10Y (8-23)

where,
·158· Petroleum Production Technology

Y = −2.69851 + 0.15841X1 − 0.55100 X12 + 0.54785 X13 − 0.12195 X14 (8-24)

and

X1 = lg ( N L + 3) (8-25)

Step 2: Obtain liquid holdup.

HL
= −0.10307 + 0.61777 ( lg X 2 + 6 ) − 0.63295 ( lg X 2 + 6 )
2
ψ (8-26)
+ 0.29598 ( lg X 2 + 6 ) − 0.0401( lg X 2 + 6 )
3 4

where,

N vL P 0.1CN L
X2 = 0.575 0.1
N vG Pa N D

and

ψ = 0.91163 − 4.82176 X 3 + 1232.25 X 32 − 22253.6 X 33 + 116174.3 X 34


(8-27)
= 0.91163 − 4.82176 X 3 + 1232.25 X 32 − 22253.6 X 33 + 116174.3 X 34

where,

N vG N L0.38
X3 = (8-28)
N D2.14

Please note ψ = 1 if X3 is less than 0.01.


Step 3: Obtain Reynolds number and friction factor. Note the friction factor can be
calculated with Chen equation or with chart.
Step4: Obtain pressure loss gradient. Modified Hagedorn-Brown model is used to
calculate the pressure loss in an example below.
Example Oil and gas flow simultaneously in a vertical well. The well does not
produce water. At a certain location in the tubing, the pressure is 1700 psig. Calculate
the pressure loss gradient at this location with Hagedorn-Brown method. Oil
production rate = 10000 stb/d ; oil density at surface = 54.7 lbm/ft3 ; oil formation
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·159·

volume factor = 1.2 bbl/stb ; oil viscosity in situ = 0.97 cP ; gas production rate = 10
million scf/d ; gas density in situ = 5.88 lb/ft3 ; gas specific gravity = 0.65 ; gas
formation volume factor = 0.009 ft3/scf ; gas viscosity in situ = 0.016 cP ; oil-water
interfacial tension = 8.41 dyn/cm ; solution GOR = 280 scf/stb ; tubing ID = 6 in ;
tubing roughness = 0.00005 ft ;
Solution Step 1: Calculate superficial velocities.
Pipe cross sectional area:
Ap = 0.25πD2 = 0.196 (ft2)
Oil in situ flow rate:
qo = 10000×1.2×5.614/86400 = 0.778 (ft3/s)
Oil superficial velocity:
vSL = 0.778/0.196 = 3.97 (ft/s)
Oil in situ density:
ρo = (54.7+0.0136×280×0.65)/1.2 = 47.61
Gas in situ flow rate:
qg = 107–10000×280×0.009/86400 = 0.757 (ft3/s)
Gas superficial velocity:
vSG = 0.757/0.196 = 3.86 (ft/s)
Mixture velocity:
vm = 3.97+3.86 = 7.83 (ft/s)
Step 2: Calculate dimensionless numbers.
4 4
ρL 47.61
N LV = 1.938vSL = 1.938 × 3.97 × = 11.87
σ 8.41
4 4
ρL 47.61
N GV = 1.938vSG = 1.938 × 3.86 × = 11.54
σ 8.41

ρL 6 47.61
N D = 120.872 D = 120.872 × × = 143.8
σ 12 8.41
4 4
1 1
N L = 0.157 μL 3
= 0.157 × 0.97 × 3
= 0.012
σ ρL 8.41 × 47.61
·160· Petroleum Production Technology

Step 3: Obtain liquid holdup.

X1 = 0.479

Y = –2.695

CNL = 0.002

X2 = 6.5×10–5

HL
= 0.267
ψ

X3=5×10–5, therefore ψ =1 and HL=0.267.

The no-slippage liquid holdup :

qL 0.778
λL = = = 0.507
qL + qg 0.778 + 0.757

It is impossible for the real holdup to be less than the no-slippage holdup,
therefore the real liquid holdup is regarded as 0.507.
Step 4: Obtain Reynolds number and friction factor.
Mixture density:

ρs = ρL H L + ρG (1 − H L ) = 47.61 × 0.507 + 5.88 × (1 − 0.507 ) = 27 (lbm ft 3 )

Mixture viscosity:

μs = μLH L μG(1− H L ) = 0.970.507 × 0.016(1−0.507) = 0.13 (cP)

Reynolds number:

ρs vm D 27 × 7.83 × 0.5
Re = 1488 = 1488 = 1.2 × 106
μs 0.13

Relative roughness:
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·161·

ε 0.00005
δ= = = 0.0001
D 0.5

According to equation or chart,

Moody friction factor = 0.013

Step 5: Calculate pressure loss gradient (ignore kinetic loss).


Elevation loss :

ΔPe 27
= ρg = = 0.188(psi ft)
Δh 144

Friction loss :

ΔPf f m ρv 2 0.013 × 27 × 7.832


= = = 0.0046(psi ft)
L 2 Dg 2 × 0.5 × 32.174

The total pressure loss gradient = 0.188+0.0046 = 0.1926 psi/ft

8.4 Tubing performance relationship

The ultimate purpose of multiphase calculations is to generate tubing performance


relationship (TPR, also know as vertical lift performance) for the well of interest.
Nowadays, multiphase flow calculations are carried out by industrial software, such as
Prosper and PipeSim. The normal process is to assume a production rate and a
wellhead pressure, the fluid pressure calculation starts from the wellhead and proceeds
downwards in steps of 100 ft for instance (or less if there are large changes in
inclination, pressure prediction, etc.).
At the same time, heat transfer is predicted and changes in fluid temperature
calculated. The process is continued until the bottom-hole pressure and temperature
have been calculated. If the predicted bottom-hole temperature does not match the real
bottom-hole temperature (usually assumed as the reservoir temperature), iterations
will be required to get a match. In essence, the pressure is predicted downwards, whilst
the temperature is predicted upwards. The process is repeated many times with
assumed flow rates. At last, the relationship between flow rate and bottom-hole
pressure can be plotted.
·162· Petroleum Production Technology

An example plot with various correlations, as calculated by the software Prosper,


is shown in Fig. 8-6. The oil has a gravity of 35ºAPI, and GOR of 800 scf/stb. Tubing
size is 4.5 in, and wellhead pressure is set at 250 psi. It can be seen there exists a
minimum bottom-hole pressure on TPR curves. This is the minimum pressure required
for the well to produce. On the right lies the friction-dominated region, where fluid
flow is stable. On the left lies the slippage-dominated region, where flow is unstable
due to severe slugging.

Fig. 8-6 TPR curves generated by multiphase flow models

The TPR curve is affected by a few production parameters, for example water-cut,
tubing size, GOR etc. Some scenarios are discussed here. ① Water-cut usually
increases after some years of production. The mixture density will increase and a
higher bottom-hole pressure will be required to lift the produced fluid. As a result, the
TPR shifts upwards. ② If a smaller tubing is installed, the friction pressure loss
increases. Therefore a higher bottom-hole pressure is needed to lift the fluids, and the
TPR curve shifts upwards. ③ If GOR increases, the mixture density decreases and a
lower bottom-hole pressure is required. Thus the TPR curve shifts downwards.
However, if the GOR is extremely high to reach annular flow regime, friction pressure
loss becomes excessive and a higher bottom-hole pressure is needed. Therefore TPR
curve shifts upwards. These effects are illustrated in Fig. 8-7.
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·163·

Fig. 8-7 TPR curves under influences of various factors

8.5 Nodal analysis for production system

The system analysis for determination of fluid production rate and pressure at a
specified node is called nodal analysis in petroleum engineering. Nodal analysis is
performed on the principle of pressure continuity, that is, there is only one unique
pressure value at a given node regardless of whether the pressure is evaluated from the
performance of upstream equipment or downstream equipment. The principle of nodal
analysis can be expressed as:

Pr – ΔP (upstream) = pressure at node (8-29)

Psep + ΔP (downstream) = pressure at node (8-30)

where,
Pr=reservoir pressure, psia;
Psep=separator pressure, psia.
The pressure loss in any component varies with flow rate. Therefore, a plot of
node pressure versus flow rate will produce two curves. The performance curve
(pressure-rate relation) for upstream equipment is the inflow performance curve. The
performance curve for downstream equipment is the outflow performance curve. The
·164· Petroleum Production Technology

intersection of the two performance curves defines the operating point, that is, the
operating flow rate and pressure, at the specified node.
For the convenience of using pressure data measured normally at either the
bottom-hole or the wellhead, nodal analysis is usually conducted using the bottom-hole
or wellhead as the solution node. This section illustrates the principle of nodal analysis
with single-diameter tubing strings.
When the bottom-hole is used as a solution node in nodal analysis, the inflow
performance is the well inflow performance relationship (IPR), and the outflow
performance is the tubing performance relationship (TPR), if the tubing shoe is set to
the top of the pay zone. Well IPR can be established with Vogel or Fetkovich method
presented in the previous chapter. TPR can be obtained with various multiphase flow
correlations discussed earlier. An example is given in Fig. 8-8. The reservoir pressure is
4000 psig at present. With the 4in tubing installed, the well produces 16000 stb/d of
liquid at a bottom-hole pressure of 3000 psig.

Fig. 8-8 An example of nodal analysis

Nodal analysis is a powerful tool for production engineers. The influences of


various factors that impact production rate can be analyzed with this approach. For
instance, the influence of tubing size is illustrated in Fig. 8-9. If 5in tubing is used, the
production rate increases to 24000 stb/d. If 3in tubing is installed, the production rate
declines to 9000 stb/d. The production rate will further increase if 6in tubing is
deployed, however the incremental is not as significant.
Chapter 8 Fluid Flow in Wells and Nodal Analysis ·165·

Fig. 8-9 Nodal analysis on tubing sizes

After some years of production, reservoir pressure declines. We can use Nodal
analysis to investigate the effect of pressure depletion on production rate.
According to Fig. 8-10, production rate drops to 9000 stb/d when the reservoir
pressure declines to 3000 psig. If the reservoir pressure further declines to 2000 psig,
the well will cease to produce and there is no cross point between IPR and TPR.

Fig. 8-10 Nodal analysis on reservoir pressure

The influence of water-cut is presented in Fig. 8-11. When water-cut increases, the
density of produced fluids also increases. Therefore, a higher bottom-hole pressure is
required to lift fluids to surface. Thus the TPR shifts upwards and the production rate
declines.
·166· Petroleum Production Technology

Fig. 8-11 Nodal analysis on water-cut

The effect of GOR on production rate is shown in Fig. 8-12. When GOR increases,
production rate first increases, but decreases at very high GOR. When gas rate is
excessive, flow regime transforms into annular flow and friction pressure loss becomes
very high. Therefore, a optimal GOR exists for the highest liquid rate. This is the
foundation for gas lift deign which will be introduced in chapter 9.

Fig. 8-12 Nodal analysis on gas rate


Chapter 9 Artificial Lift Methods
When the field ages, the reservoir pressure declines and the production rate drops.
Artificial lift methods are often employed to boost or maintain production. Common
artificial lift methods include sucker rod pump, gas lift and electrical submersible pump
(ESP). Other types of artificial lifts are also briefly introduced.

9.1 Sucker rod pump

Sucker rod pumping is also referred to as beam pumping. Beam pumps are very
common on land wells, but are usually limited to low production rates of a few hundred
barrels per day. They use mechanical energy to lift oil from well to surface. Rod pumps
are simple and easy to operate. They can pump a well at very low pressure to maximize
oil production rate. They are applicable to slim holes, multiple completions, and
high-temperature and viscous oils. The major disadvantages of beam pumping include
excessive friction in deviated wells, solid-sensitive problems, low efficiency in gassy
wells, limited depth due to rod capacity, and bulky in offshore operations. A sucker rod
pumping system consists of a pumping unit at surface and a plunger pump submerged
in the production liquid in the well, as seen in Fig. 9-1.
The prime mover is either an electric motor or an internal combustion engine.
Electric motors are most common because they can be easily automated. The power
from the prime mover is transmitted to the gear reducer by a V-belt. The gear reducer
drives the crank arm at a lower speed of 4~40 r/min depending on well characteristics
and fluid properties. The rotary motion of the crank arm is converted to an oscillatory
motion by the walking beam through a pitman arm.
Counterbalance for conventional pumping units is accomplished by placing weights
directly on the beam or by attaching weights to the rotating crank arm. The horse head and
the hanger cable arrangement is used to ensure that the upward pull on the sucker rod string
is always vertical at all times. The polished rod and stuffing box combine to maintain a
good liquid seal at the surface, thus forcing fluid to flow into the connection just below the
stuffing box. Conventional pumping units are available in a wide range of sizes, with stroke
lengths varying from 12 to almost 200 in. Walking beam ratings are expressed in allowable
·168· Petroleum Production Technology

polished rod load (PRL) and vary from approximately 3000 to 35000 lb.

Fig. 9-1 Configuration of sucker rod pump

The American Petroleum Institute (API) has established designations for sucker rod
pumping units using a string of characters containing four fields, for example,
C-228D-200-74. The first letter stands for the type of pumping unit. A is for air-balanced
units (Fig. 9-2), B for beam counterbalance units, C for conventional units (Fig. 9-3),

Fig. 9-2 Air-balanced pumping unit


Chapter 9 Artificial Lift Methods ·169·

Fig. 9-3 Conventional pumping unit

Fig. 9-4 Lufkin Mark II unit

and M for Mark II units (Fig. 9-4). The number 228 is the peak torque rating in thousands
of inch-pounds. D stands for double-reduction gear reducer. The third field is the code for
PRL rating in hundreds of pounds. The last field is the code for stroke length in inches.
The working cycle of a plunger pump is illustrated in Fig. 9-5. The pump is
installed in the tubing below the dynamic liquid level. It consists of a working barrel
and liner, standing valve, and traveling valve at the bottom of the plunger. As the
·170· Petroleum Production Technology

plunger moves downward, the traveling valve opens and allows the fluid to pass
through the valve. The standing valve is closed during this downward motion of the
plunger. When the plunger is at the bottom of the stroke and starts an upward stroke,
the traveling valve closes and the standing valve opens. As upward motion continues,
the fluid above the traveling valve is lifted to surface. At the mean time, fluid passes
the standing valve and enters the chamber. The fluid continues to fill the chamber
above the standing valve until the plunger reaches the top of its stroke.

Fig. 9-5 A pumping cycle

Different plunger sizes in both area and length allow for a wide range of possible
flow rates. For a given plunger size, the flow rate can be further adjusted by altering
stroke length and pump speed. Even lower flow rates can easily be accommodated by
cycling the pump on and off. Finding the right balance between stroke length and pump
speed is the art of beam pump design.
Pump performance is analyzed with surface dynamometer card. The surface dyna-
mometer is a device that records the motion of the polished rod during the pumping cycle.
The rod string is forced by the pumping unit to follow a regular time versus position pattern.
However, the polished rod reacts with the loadings that are imposed by the well. The
surface dynamometer cards record the history of the variations in loading on the polished
rod during a cycle. The purpose of the card is to check pump conditions and diagnose pump
problems. The dynamometer measures the force on the polished rod and the distance that
the rod moves. Force multiplying distance equals work. The dynamometer thus records the
work done during one cycle. A typical example is given in Fig. 9-6.
Chapter 9 Artificial Lift Methods ·171·

Fig. 9-6 Ideal dynamometer graph

Many issues can affect the performance of rod pumps, and these issues can be
diagnosed with dynamometer card. These common issues include liquid pound, well
pumping off, improper counterbalance, sand accumulation inside pump, scale and wax
deposition inside pump. Fluid pound occurs when insufficient fluid enters the pump
barrel on the pump upstroke. A space filled with gas exists above the liquid. During the
down-stroke, the traveling valve does not open until the plunger hits the liquid surface.
This impact sends shock waves up the rod to surface and shows on the dynamometer card.
An example is given in Fig. 9-7. This graph indicates the fluid supply is insufficient. The
pump should slow down, or be replaced with a smaller-diameter pump.

Fig. 9-7 Liquid pound effect on dynamometer graph


·172· Petroleum Production Technology

9.2 Gas lift

Gas lift improves production by injection of compressed gas into tubing through
annulus and valves installed on tubing string. The compressed gas boosts production by
two means. Firstly, gas expansion propels oil to flow to surface. Secondly, gas
decreases the density of the fluid column in the well. The depleted well flowing
pressure can thus support the flow of gas-liquid mixture.
Gas lift method has been widely used in oil fields. Gas lift does not employ a
pump, therefore it can not yield very low bottom-hole pressure. It can be used in sandy
or deviated wells in onshore or offshore fields. Well depth is not a problem. Lifting
costs for a large number of wells are generally low if gas compression is required.
However, source of lift gas nearby is required.

9.2.1 Gas lift system


A complete gas lifting system includes a gas compressor, injection manifold, gas-liquid
separator, and tubing string with gas lift valves installed. The surface equipment is
shown in Fig. 9-8 and the down-hole configuration is shown in Fig. 9-9. It can be seen
multiple gas lift valves are installed, which can reduce the pressure requirement on
gas compressors to kick off the well flow. Besides, installation of multiple valves can

Fig. 9-8 Surface equipment for gas lift


Chapter 9 Artificial Lift Methods ·173·

Fig. 9-9 Down-hole configuration for a gas lifted well

achieve deeper injection depths with the available injection pressure. In Fig. 9-9, PWH
represents the well head pressure, Pcs the pressure of gas for injection, Gu the mixture
gradient, and GR the gradient of formation fluid.
Gas lift can be continuous or intermittent. For continuous gas lift, gas is injected
continuously into the tubing. Continuous gas lift is used in wells with good
productivity above 0.5 (stb/d)/psi, plus relatively high reservoir pressure and good
supply of liquids. For intermittent gas lift, gas is injected in a start-and-stop fashion,
which is suitable for wells with low liquid supply.
The most common gas lift valves are pressure-controlled valves. The valve
opening can be controlled by casing pressure or tubing pressure. The dome pressure
and the springs tend to close the valve, while the pressure in casing or tubing tends to
open the valve, as presented in Fig. 9-10.
·174· Petroleum Production Technology

Fig. 9-10 Pressure-controlled gas lift valves

9.2.2 Well unloading and kick-off


Before gas injection is initiated, the tubing is filled with reservoir liquids (mostly water
in reality), while the casing is filled with packer fluid, completion fluid, or drilling
mud. The injected gas must propel the packer fluid into the tubing, moreover the gas
pressure must be high enough to overcome the fluid hydrostatic pressure in tubing.
An example of continuous gas lift is given in Fig. 9-11. Both tubing pressure and
casing pressure are plotted in the figure. Based on the plot, one unloading valve should
be installed at the true vertical depth (TVD) of 11000 ft. A gas injection pressure of
3500 psig is required at wellhead to kick off the well. But such a simple system does
require a high gas pressure to start the well flow when the column of liquid inside the
completion is dense. This pressure may be too high for the surface gas compressor to
achieve. Therefore, a dedicated kick-off compressor can be installed to supply such a
high pressure.
In practice, a more common solution is to install multiple unloading valves
situated in gas lift mandrels distributed down the tubing string. These valves are
initially open and allow shallow gas lift. As the column of liquid in the completion is
lightened, and the pressure in either the tubing or the casing reduces, the valves
automatically close, pushing the lift gas progressively deeper down the well.
Chapter 9 Artificial Lift Methods ·175·

Fig. 9-11 Well kick-off with single unloading valve

Gas lift design requires two aspects: determination of gas injection rate and locations
for unloading valves. The optimum gas injection rate can be obtained with pressure
traverse curves or nodal analysis. An example of nodal analysis is given in Fig. 9-12.
While the GOR or GLR (gas liquid ratio) increases, the liquid rate also increases
initially. But liquid rate begins to decline when an optimum GLR is exceeded. For
example, the GOR of a crude oil is 100 scf/stb. The nodal analysis in Fig. 9-12 shows
the optimum GLR of 1000 scf/stb yields a liquid rate of 15000 stb/d. The water-cut of
produced liquid is at 60%. Therefore, the gas injection rate should be :
15000×1000–15000×(1–0.6)×100 = 14.4×106 scf/d.
Another example with traverse curves is given in Fig. 9-13. Two conclusions can
be drawn based on the chart. ①Deeper injection point is more effective at reducing the
well flowing pressure. ②Optimum gas injection rate can be determined based on the
curves. For the deep gas injection, the well flowing pressure still reduces at high gas
injection rates, but the effect is not as significant.
·176· Petroleum Production Technology

Fig. 9-12 Nodal analysis on effect of GLR

Fig. 9-13 Effect of gas injection rate on pressure traverse curves

The traverse curve for the optimum gas injection rate is thus the optimum
traverse curve applied to spacing of unloading valves, as long as the compressors are
able to generate the optimum gas flow rate. An example is given in Fig. 9-14. The
available gas injection pressure at wellhead is only 1500 psig. Based on the gas
pressure curve and the tubing pressure curve, the first valve is placed at 5400 ft TVD.
A pressure margin is applied to account for the pressure loss across the valve, usually
100 to 200 psig.
Chapter 9 Artificial Lift Methods ·177·

Fig. 9-14 Location of the first unloading valve

Fig. 9-15 shows how to locate the second valve. From the tubing traverse curve at
the depth of the first valve, draw a line parallel to the formation fluid gradient and find
the location for the second valve. Again, we need to give a pressure margin for pressure
loss in the valve. When the second valve opens, the first valve senses the pressure
reduction in tubing and closes.

Fig. 9-15 Location of the second unloading valve


·178· Petroleum Production Technology

This procedure is repeated until the depth of packer is reached, as shown in Fig.
9-16. Compare with installing only one unloading valve, the required wellhead gas
pressure is now reduced to 1500 psig.

Fig. 9-16 Complete design for valve spacing

In reality, the pressure in tubing often changes and it is difficult to predict tubing
pressures. The pressure in casing is always known. Therefore it is actually easier to
design gas lift with casing-pressure operated valves. A sample design is given in Fig.
9-17. This design will lead to shallower injection depth than the previous design with
tubing-pressure operated valves.
The performance of a gas lift system is monitored by observing injection and
produced fluid flow rates and the casing and tubing pressures. Diagnostic surveys such
as production logging can also be carried out to determine if any valve is not working
properly. Optimizing the amount of gas going to each well is also critical. With too
much lift gas, the system will underperform due to increased friction. Too little gas and
the reservoir fluids will not be lightened enough.
An alternative to continuous gas lift is intermittent gas lift. The equipment for
intermittent gas lift is similar to that needed for continuous gas lift, but the operating
principle is different. For the continuous gas lift system, the gas phase is dispersed in
the liquid phase, while intermittent gas lift relies on a finite volume of gas lifting a
liquid column to surface at regular intervals.
Chapter 9 Artificial Lift Methods ·179·

Fig. 9-17 Spacing of casing-pressure operated valves

9.3 Electrical submersible pump

ESPs are widely used in oil fields, especially for high production rate wells. The
arrangement for EPS is shown in Fig. 9-18. The power cable supplies electricity for the
motor, and the motor drives the pump. The pump is made of many stages of impeller-

Fig. 9-18 Configuration of ESP


·180· Petroleum Production Technology

Fig. 9-19 Impeller and diffuser

diffuser pairs, as presented in Fig. 9-19. ESP can lift very high volumes with high
efficiency. They can be used in onshore and offshore wells. But ESP requires high
voltage electricity. Gas and sand production can be troublesome to ESP.
A centrifugal pump performs work by lifting the fluid a certain head. The head is
the vertical distance the fluid is lifted. The head is independent of the fluid. Fig. 9-20
shows the output of one stage only. The total required head determines the number of
stages and the power requirement of motor. ESP design concerns itself primarily with

Fig. 9-20 Head of liquid produced by one stage of ESP


Chapter 9 Artificial Lift Methods ·181·

choosing the right type of pump, the optimum number of stages, and the corresponding
motor and cable size to ensure the smooth functioning of the system. As can be seen
from the diagram, changes in well productivity are difficult to accommodate.
The performance of the system is monitored primarily by the use of a current and
voltage meter, measuring the motor load and by analyzing the fluid throughput against
the hydraulic head. If the rates are too high or too low for the pump, then significant
thrust loads will develop on the pump stages and the pump may self-destruct. Other
problems include electrical short circuits, especially when cable penetrators are
required through packers.
ESPs are classified by their outside diameters ranging from 3.5 to 10 in. The
number of stages is determined by the volumetric flow rate and the pressure
requirement. Run lives vary enormously from many years in shallow, low temperature,
solid-free wells to less than a year in more extreme environments.
Example: You are required to design an ESP to produce 8000 stb/d against a
wellhead pressure of 100 psia. The available information is given below. Oil gravity =
32° API ; oil FVF = 1.25 ; oil viscosity = 5 cP ; GOR = 50 scf/stb ; water-cut = 0 ; gas
specific gravity = 0.7 ; Reservoir pressure = 4350 psia ; AOF = 15000 stb/d ; Well
depth = 10000 ft ; Tubing ID = 2.992 in ; Casing OD = 7 in.
(1) The ESP is required to deliver a higher rate at depth due to shrinkage of liquid.

QESP = 8000×1.25 = 10000 (stb/d)

Select an ESP that delivers the design rate around its maximum efficiency. An example
chart is given in Fig. 9-21.

Fig. 9-21 An example performance chart for 100-stage ESP


·182· Petroleum Production Technology

(2) Well IPR gives the well flowing pressure to support the design flow rate.
2
Qo P ⎛P ⎞
= 1 − 0.2 wf − 0.8 ⎜ wf ⎟
AOF Pr ⎝ Pr ⎠

We can calculate Pwf = 2823 psia


(3) The pump can be placed at 9800 ft. The pump suction pressure is:
Ps = 2823 – 0.375×200 = 2748 (psia)
(4) With multiphase flow models, compute the required pressure at pump
discharge.
Pd = 3728 psia
(5) The pressure differential is
Pd – Ps = 3728 – 2748 = 980 (psi)

The required pumping head is:


Head = 980/0.433 = 2263 (ft of freshwater)
The chart shows 100 stages can pump 6000 ft (60 ft per stage). Therefore,
Required stages = 2263/60 = 38 (stages)
The chart shows 100 stages can produce horsepower of 600 hp①(6 hp per stage).
Therefore,
Horsepower requirement = 38×6 = 228 (hp)

9.4 Other artificial lift methods


Hydraulic submersible pumps (HSP) operate in a way similar to ESP, using multi-stage
centrifugal pumps. But a down-hole turbine powers the pump. The HSP configuration is
presented in Fig. 9-22. HSP operates at a higher speed than an ESP (around three to four
times higher revolutions/min). HSP therefore requires fewer stages and is smaller. The
turbine requires no electrical connections or down-hole electronics. But it does require a
power fluid to be pumped down-hole. This power fluid can be commingled with the
reservoir fluid and returned to the surface, or it can be returned in a separate conduit or
disposed of down-hole.
————————
① 1hp=745.7W.
Chapter 9 Artificial Lift Methods ·183·

Fig. 9-22 Configuration of HSP

Jet pumps require no down-hole moving parts. The jet pump configuration is
presented in Fig. 9-23. They find wide applications generally in wells of low to
moderate rates. The technology has been around for centuries and is found in many
surface oil and gas applications – anywhere where high-pressure fluids can be used to
boost a low-pressure fluid. The power fluid and the reservoir fluid must mix, so a key
issue is the selection of an appropriate power fluid. Jet pumps are compact and reliable,
and can be easily installed and retrieved by wire line. But jet pump is less efficient than
other pump systems and requires large volumes of power fluids.
·184· Petroleum Production Technology

Fig. 9-23 Configuration of jet pump

Progressive cavity pumps (PCP) are positive displacement pumps in nature.


The pump configuration is shown in Fig. 9-24. Their operation involves the rotation
of a metal spiral rotor inside either a metal or an elastomeric spiral stator, as shown in
Fig. 9-25. Rotation causes the displacement of a constant volume cavity formed by
the rotor and the stator. The area and the axial speed of this cavity determine the
production rate.
Chapter 9 Artificial Lift Methods ·185·

Fig. 9-24 Configuration of PCP

Fig. 9-25 Rotor and stator


·186· Petroleum Production Technology

9.5 Discussions

Sucker rod pumps, ESP and gas lift are the dominant artificial lift methods in the
market. While selecting an artificial lift method, we need to consider a few factors. Tab.
9-1 provides an overview of the application envelope and the respective advantages
and disadvantages of the various artificial lift techniques. Gas lift and ESP are
recommended for high flow rates and offshore applications. If the well produces large
quantities of solids, gas lift is the suitable method because gas lift does not use a pump,
thus eliminating erosion problem. Beam pumps are generally unsuited to offshore
applications because of the low rate and the large area they occupy. Even though most
of the world’s artificially lifted wells are beam pumped, the majority of these wells are
stripper wells producing less than 10 bbl/d.
Tab. 9-1 Comparison of artificial lift methods
  Offshore application Solid handling  Maximum temperature/℃  Volume capacity/(m3/d) 

Rod pump  Not recommended Less than 200 ppm 260 Less than 100 
Gas lift  Excellent  Excellent 175 8-8000
ESP  Excellent  Less than 200 ppm 160 40-8000
PCP  Good  Up to 5% 120 5-800
HSP  Good  Less than 200 ppm 260 15-800
Jet pump  Good  Less than 200 ppm 260 15-2500
Chapter 10 Acidizing and Fracturing
During drilling and production operations, a few factors may damage the permeability
of near-well region. This phenomenon is named formation damage. As a result,
formation damage leads to reduction in well production rate. In practice, acids are
routinely injected into reservoirs with an attempt to restore reservoir’s original
permeability. Besides, fracturing operations are routinely carried out to create fractures
that extend into the reservoir, bypassing the damaged zone.

10.1 Formation damage and skin factor

Many factors can result in formation damage. Common causes include particle
plugging, solid precipitation, and perforating operation. Fracturing and acidizing
operations are routinely conducted to remediate formation damage and improve well
productivity. However, the induced damage may or may not be repairable.
(1) Plugging of pores by particles is a very common cause of damage. Drilling
mud contains high concentration of particles. Completion fluid and the injected water
(for water flooding) also contain certain concentration of particles. When the small
particles enter the rock, they may deposit inside the pore throats, reducing permeability
significantly. Besides, large particles may also bridge at surface of rock and form a
filter cake. For example, particles in drilling mud form a mud cake at sand face. On the
other hand, when the salinity of formation water changes, the internal clay mineral in
reservoir rock begins to disperse or swell, thus reducing original rock permeability.
Bacteria growth in reservoir rock is another factor that reduces permeability.
(2) Scale precipitation and asphaltene precipitation inside reservoir rock also
reduce rock permeability. Inorganic precipitates include calcium or barium, combined
with carbonate or sulphate ions. The ions in reservoir water were initially in a chemical
equilibrium. The injected water alters this equilibrium and yields scale problem. For
example, the calcium ions in injected water yields calcium carbonate precipitates. On
the other hand, pressure and chemical changes in reservoir lead to asphaltene
precipitation.
(3) During perforating process, the fired charges yield a crushed and compacted
·188· Petroleum Production Technology

zone around the perforations. It was reported the rock permeability near the perforations
is reduced by 80%. While the fluids converge into the well, the reduction in permeability
causes a high resistance to flow.
Formation damage leads to extra pressure loss in the damaged zone, which is
referred to as skin effect. Hawkins developed a formula to describe skin effect. It is
assumed that the damaged zone has a uniform permeability that extends a certain
radius into the reservoir, as depicted in Fig. 10-1.

Fig. 10-1 Damaged zone and skin effect

The damage induces an additional pressure loss that is expressed as:

ΔPs = ΔP ( in skin zone due to ks ) − ΔP ( in skin zone due to k ) (10-1)

Applying the Darcy’s law yields:

Qo Bo μo r ⎛ Q B μ ⎞ r
ΔPs = ln s − ⎜ o o o ⎟ ln s (10-2)
0.00708hks rw ⎝ 0.00708hk ⎠ rw

Rearrange and obtain the expression below,

Qo Bo μo ⎛ k ⎞ rs
ΔPs = ⎜ − 1⎟ ln (10-3)
0.00708kh ⎝ ks ⎠ rw

where,
ΔPs = pressure loss induced by damage, psi;
Chapter 10 Acidizing and Fracturing ·189·

Qo = oil flow rate in reservoir, stb/d;


k = original reservoir permeability, mD;
ks = permeability of the damaged zone, mD;
rs = radius of damaged zone, ft;
h = reservoir thickness, ft;
βo = oil formation volume factor, bbl/stb;
μo = oil viscosity, cP.
The skin factor is expressed as:

⎛k ⎞ r
S = ⎜ − 1⎟ ln s (10-4)
⎝ ks ⎠ rw

The additional pressure loss is thus:

⎛ Q B μ ⎞ QB μ
ΔPs = ⎜ o o o ⎟ S = 141.2 o o o S (10-5)
⎝ 0.00708kh ⎠ kh

Skin factor quantifies the severity of formation damage. A positive skin factor
indicates the near well region has been damaged. A negative skin factor is a result of
well stimulation or horizontal well. A zero skin factor indicates no change occurs to the
reservoir permeability. For fluid flow under steady state in reservoir,
Actual pressure loss = ideal pressure loss + damage – induced pressure loss
Applying Darcy’s law obtains:

⎛ Q B μ ⎞ r ⎛ Q B μ ⎞
Pi − Pwf = ⎜ o o o ⎟ ln e + ⎜ o o o ⎟ S (10-6)
⎝ 0.00708kh ⎠ rw ⎝ 0.00708kh ⎠

Rearrange to obtain the flow rate that takes into account the skin factor:

0.00708kh ( Pi − Pwf )
Qo = (10-7)
⎛ r ⎞
μo Bo ⎜ ln e + S ⎟
⎝ rw ⎠

where,
S = skin factor;
Pi = initial reservoir pressure, psi;
Pwf = well bottom hole pressure, psi.
·190· Petroleum Production Technology

10.2 Fracturing

It is a common practice to fracture the reservoir rock to bypass damaged zone and
improve productivity. In North America, around 50% of wells are fractured at
completion stage. Many low-permeability reservoirs have to be fractured to produce
economically. The principal of fracturing is to pump fluid down the well at a pressure
high enough to break down reservoir rock. Then proppants are injected into fractures to
prevent closures of fractures, thus maintaining high productivity.

10.2.1 Fracture pressure


Rock fracture pressure is determined by the tensile strength of rock. The tensile
strength is defined as the pulling force that is required to rupture a rock sample, divided
by the cross sectional area of the rock sample. The tensile strength of rock is only 10%
compressive strength. Thus, a rock is more likely to fail in tension than in compression.
Let’s look at a rock sample buried underground, as shown in Fig. 10-2. Three
perpendicular stresses apply to the rock sample. The vertical stress is the overburden
stress, which is normally much higher than the horizontal stresses. The other two
horizontal stresses directly influence the fracturing of rock. In theory, the pressure
required to rapture a rock should be greater than the minimum principal stress. Based
on the above explanation, we can understand the fact that most fractures are created in
the vertical direction.

Fig. 10-2 Three principal stresses

Fracture pressure can be directly obtained by leak-off test, as illustrated in Fig.


10-3. ① Fill the well with mud while the drill string is suspended in the well.
Chapter 10 Acidizing and Fracturing ·191·

1
② Close the annular BOP. ③ Pump the mud in small increment, for instance
4
barrel. ④ Plot injection pressure versus injected volume or time.

Fig. 10-3 Leak-off testing equipment

The injection pressure will first build up linearly when mud is injected. The
pressure will become nonlinear afterwards, when rock starts to break down. At point B
in Fig. 10-4, the pressure will suddenly drop, which indicates a fracture has been
created. Afterwards, fracture propagates to deep formation at lower pressure. Note that
we need to add mud hydrostatic pressure to pump injection pressure to obtain the
fracture pressure.

Fig. 10-4 Result of a typical leak-off test


·192· Petroleum Production Technology

Fracture pressure can also be predicted with various rock mechanics models. Two
methods are introduced here.
Hubbert and Willis method:
1 2
FP = OS + PP (10-8)
3 3
Eaton method:
PR
FP = ( OS − PP ) + PP (10-9)
1 − PR
where,
FP = Fracture pressure;
OS = Overburden stress;
PP = Pore pressure;
PR = Poisson’s ratio.

10.2.2 Fracturing fluid and proppants


The fracturing operation consists of several stages, as illustrated in Fig. 10-5. Firstly, a
fluid without proppants is injected to initiate fractures. This fluid is often referred to as
pad. After the pad injection, proppant slurry at a low proppant concentration is added to
the fracturing fluid and pumped into the well. The concentration of proppants increases

Fig. 10-5 Injection of fracturing fluid and proppants


Chapter 10 Acidizing and Fracturing ·193·

gradually to a designed level. When the proppants reach the front tip of the fracture, tip
screen out occurs and the fracture ceases to grow in length. Afterwards, a flush fluid is
injected to displace the proppants inside the wellbore into the fractures.
A typical water-based fracturing fluid consists of water and a thickening polymer
such as hydroxypropyl guar (HPG). The concentration of HPG is often between 20 to
80 lb/1000 gallon to achieve desirable viscosity. Ideally, the fracturing fluid
demonstrates high viscosity under low shear rate, but low viscosity under high shear.
This feature yields low pressure loss in tubing. While in fractures, the viscous fluid
produces wider fractures and provides better transportation of proppants. Other types of
fracturing fluids include oil-based fluid and emulsion.
Past experiences show the fracturing fluid should have a viscosity higher than
100 cP at 170/s to effectively transport proppants. It is very common to add crosslink
agents to fracturing fluid to boost its viscosity. Crosslink agents include borate,
titanate and zirconate. They form bonds with polymer chains, resulting in very large
polymer networks and very high viscosity. Borate can be used up to 225ºF and the
other two agents can function up to 350ºF. Nowadays, crosslink agents in capsules
are mixed with fracturing fluid, so that the delayed crosslink effect occurs inside the
fracture.
After the injection is complete, the polymer fluid must break down to very low
viscosity and flow back. Otherwise, the polymer fluid plugs the proppant pack, which
leads to very low fracture permeability. Breakers serve the purpose of breaking down
the polymer immediately after the treatment. Encapsulated breakers are widely used
nowadays. They become effective by physical or chemical means immediately after the
treatment.

10.2.3 Models of fracture geometry


Two of the most widely-used 2D models for fracture geometry design are PKN model
and KGD model. The Perkins Kern Nordgren (PKN) model (Fig. 10-6) is usually
applied when the fracture length is much greater than the fracture height, while the
Khristianovic Geertsma deKlerk (KGD) geometry (Fig. 10-7) is used if fracture
height is more than the fracture length. In certain formations, either of these two
models can be used successfully to design hydraulic fractures.
·194· Petroleum Production Technology

Fig. 10-6 Geometry of PKN model


L = fracture length ; h = fracture height ; H = maximum fracture height ; Ww = maximum fracture width

The PKN model is expressed as follows:

0.25
0.3πγ ⎡ (1 − ϑ ) Qμxf ⎤
wa = ⎢ ⎥ (10-10)
4 ⎣ G ⎦

The KGD model is expressed as follows:

0.25
0.29π ⎡ (1 − ϑ ) Qμxf ⎤
2
wa = ⎢ ⎥ (10-11)
4 ⎣⎢ Gh ⎦⎥

where,
wa = average fracture width, in;
γ = geometric factor, normally regarded as 0.75;
ϑ = Poisson ratio;
Q = injection rate, barrel/min;
μ = viscosity of fracturing fluid, cP;
xf = fracture half length, ft;
G = elastic shear modulus, psi;
h = fracture height, ft.
Chapter 10 Acidizing and Fracturing ·195·

Fig. 10-7 Geometry of KGD model

Rw = radius of wellbore;
Vx = velocity of fluid.
Both models reveal that wider fractures can be achieved by increasing injection
rate, increasing the viscosity of fracturing fluid, or increasing fracture length. Both
models indicate that to double fracture width, the injection rate or fluid viscosity have
to increase by 16 fold. This is not realistic in field operation. Therefore, the fluid rate
and viscosity have insignificant influence on facture width.
The key is to use models to make decisions, rather than trying to calculate precise
values for fracture dimensions. The design must always compare actual results with the
predictions from model calculations. By calibrating the models with field results, they
can be used to make design changes and improve the success of stimulation treatments.
If the correct fracture height value is used in a 2D model, the model will give
reasonable estimates of created fracture length and width if other parameters, such as
in-situ stress, Young’s modulus, formation permeability, and total leak-off coefficient
are also known and used.

10.3 Matrix acidizing

For a matrix acidizing operation, acid is injected into a damaged reservoir, without
fracturing the formation. In general, formation damage due to mud invasion, clay
·196· Petroleum Production Technology

swelling or fines migration can be partially or totally removed by acid. But acid can not
treat mechanical skin, such as partial penetration or perforation damage. Acid treatment
aims at removing damage in the near-well region, rather than reaching deep formations.
For sandstones, a typical acid treatment begins with a pre-flush with HCl,
followed by injection of HF and HCl mixture. A post-flush with brine or diesel
displaces the acids inside tubing or wellbore. Finally, the injected acids must flow
back to minimize damage due to precipitation. For many years, carbonates have been
treated with 15% HCl. For sandstones, a pre-flush of 15% HCl, followed by 3% HF
plus 12% HCl is standard procedures. In fact, the HF(3%)/HCL(12%) mixture is so
common that it is referred to as mud acid. The reactions between acids and minerals
are presented as follows:
Calcite:

2HCl + CaCO3 → CaCl2 + CO 2 ↑ + H 2 O

Dolomite:

4HCl + CaMg ( CO3 )2 → CaCl2 + MgCl2 + 2CO 2 ↑ +2H 2 O

Siderite:

2HCl + FeCO3 → FeCl2 + CO 2 ↑ + H 2 O

Quartz:

4HF + SiO 2 ←→ SiF4 ( silicon tetrafluoride ) ↑ +2H 2 O

SiF4 + 2HF ←→ H 2SiF6 (fluosilicic acid)

The recommended fluid types and concentrations for sandstone treatments are
presented in Tab. 10-1 and Tab. 10-2. However, field experiences and lab results are
most important in designing acid treatments.
Tab. 10-1 Recommended pre-flush fluid for sandstone (k = permeability)
Mineralogy k > 100mD 20mD < k < 100mD k < 20mD
Silt<10% and clay<10% 15% HCl 10% HCl 7.5% HCl
Silt>10% and clay>10% 10% HCl 7.5% HCl 5% HCl
Silt>10% and clay<10% 10% HCl 7.5% HCl 5% HCl
Silt<10% and clay>10% 10% HCl 7.5% HCl 5% HCl
Chapter 10 Acidizing and Fracturing ·197·

Tab. 10-2 Recommended acidizing fluid for sandstone (k = permeability)


Mineralogy k > 100mD 20mD < k < 100mD K < 20mD

Silt<10% 12% HCl 8% HCl 6% HCl


and clay<10% plus 3% HF plus 2% HF plus 1.5% HF

Silt>10% 13.5% HCl 9% HCl 4.5% HCl


and clay>10% plus 1.5% HF plus 1% HF plus 0.5% HF

Silt>10% 12% HCl 9% HCl 6% HCl


and clay<10% plus 2% HF plus 1.5% HF plus 1% HF

Silt<10% 12% HCl 9% HCl 6% HCl


and clay>10% plus 2% HF plus 1.5% HF plus 1% HF

For carbonates, weak acid is recommended for perforation cleanup. Strong acid
should be used to acidize carbonates, unless HCl causes corrosion problems which may
occur in very deep wells. Experiences have shown higher acid concentration yields
deeper penetrations. Therefore, high acid concentration is recommended for carbonate
acidizing, as shown in Tab. 10-3.
Tab. 10-3 Recommended acid for carbonate treatment
Purpose Recommended Acid

Perforating fluid 5% acetic acid

9% formic acid
Perforation cleanup 10% acetic acid
15% HCl

15% HCl
Deep damage
28% HCl

In sandstone acidizing process, the reaction rate at sand surface is slow and acid
propagates with a relatively uniform front. However for carbonates, the reaction is fast
and the front is not uniform. When acid is injected into the carbonates, the large pores
receive more acid volume than the small pores and dissolute at a much higher rate.
Afterwards the large pores grow even larger and continue to receive more acid, and
eventually grow into large channels. These large channels are named wormholes, as
shown in Fig. 10-8. Fluid flows in wormholes and bypasses the damaged region. The
fluid pressure loss in wormholes are negligible due to the large sizes compared with
normal pore sizes in carbonates. As such, it is critical to predict the depth of wormholes
which controls the skin factor after treatments.
·198· Petroleum Production Technology

Fig. 10-8 Acid induces wormholes in carbonate rock

The formation of wormhole and the structure of wormholes depend on the rates of
surface reaction, diffusion and fluid loss. At low injection rates, only large wormholes
will form with very few branches. At high injection rate, a dominant wormhole
develops together with a few braches and propagates into the rock. When the injection
rate increases further, the injected acid is consumed by the small branches and the
acidizing efficiency decreases. This indicates an optimal injection rate exists for
effective wormhole propagation.
Chapter 11 Sand Control
Sand control techniques are often necessary when the completion is required to
mechanically hold back the movement of formation sand. Common sand control
methods include gravel pack and screen. In essence, the size distribution of formation
grain is the foundation for selection and design of appropriate sand control methods.

11.1 Grain size distribution

A sandstone is formed by the deposition of sand particles of various sizes. There are
two techniques to determine size distribution of particles: Sieve analysis and laser
particle size (LPS) analysis. LPS is gaining more popularity because it is cheaper,
quicker, and requires a small sample. Before the measurement is done, the core sample
needs to be prepared following certain standards.
The core is first cleaned to remove oil and brine. Solvents such as methanol and
chloroform are used with additives to prevent damage to clay minerals. This process
can take several weeks with heavy oil reservoirs. Afterwards, the core is slowly dried
then broken up using a pestle and mortar. Care must be taken not to grind or crush
grains and this can be confirmed by a microscope. The microscope will also confirm
when the disaggregation is complete and only single grains remaining.
For LPS analysis, the sample is placed in water or non-aqueous fluid with a
dispersant to prevent aggregation. LPS uses a laser and photosensitive detector to
measure the scattering of light caused by diffraction. It can detect particles down to
0.1mm. The measurement is converted into an equivalent diameter by the assumption
that each particle is a sphere.
The results from sieve and LPS analysis are usually presented as cumulative
distributions. An example of LPS analysis is shown in Fig. 11-1. The different curves
represent samples from different depths in the same well. Raw data from a particle size
analysis usually have to be reformatted into a semi-log plot format.
·200· Petroleum Production Technology

Fig. 11-1 Examples of particle size distribution

In Fig. 11-1, the probability distribution coefficients are labeled as D%. For
example, on the most right curve, D40 equals 100 μm. This indicates 40% of particles
have diameters less than 100 μm. D50 represents the median particle diameter. From
these coefficients, various ratios can be calculated. The common ones are the D40/D90
and the D10/D95. These ratios represent the degree of sorting of the formation. The
D40/D90 is often referred to as the uniformity coefficient, while D10/D95 the sorting
coefficient.
Fig. 11-2 presents five samples. Sample No. 1 is well sorted, while No. 4 and No.
5 are poorly sorted. It can be seen sieve analysis and LPS yield different size
distributions. Another parameter commonly extracted is the volume percent of fines.
This is defined as those particles that pass through a 325 US mesh screen (i.e. particles
less than 44 mm) and approximates the size range of particles that can cause plugging
problems in both screens and gravel packs.
Another parameter often used is the mesh size, particularly for gravels and
proppants. The mesh size represents the number of openings per inch on a sieve.
Smaller numbers therefore represent coarser particles. Commercial products such
as gravels are required to cover a narrow particle size range. For example, a 14/20
gravel contains particles that can pass through a 14-mesh sieve, but not a 20-
mesh sieve.
Chapter 11 Sand Control ·201·

Fig. 11-2 Examples of well-sorted and poorly-sorted samples

In reality, wire diameters are standardized as shown in Tab. 11-1, although wear
and distortion can cause errors. Assuming a uniform volumetric distribution of particle
sizes, the mean (median) particle size of a sieved product is the mean of the gap
dimensions. Thus, a 12/20 gravel has a mean particle diameter of 1260 μm, whilst a
12/16 gravel has a mean particle size of 1435 μm and a better permeability.

Tab. 11-1 Wire mesh and gap dimensions Continued


Gap Dimension
U.S. mesh
in μm mm

3 0.2650 6730 6.730

4 0.1870 4760 4.760

5 0.1570 4000 4.000

6 0.1320 3360 3.360

7 0.1110 2830 2.830

8 0.0937 2380 2.380

10 0.0787 2000 2.000

12 0.0661 1680 1.680

14 0.0555 1410 1.410


·202· Petroleum Production Technology

Continued
Gap Dimension
U.S. mesh
in μm mm

16 0.0469 1190 1.190

18 0.0394 1000 1.000

20 0.0331 841 0.841

25 0.0280 707 0.707

30 0.0232 595 0.595

35 0.0197 500 0.500

40 0.0165 400 0.400

45 0.0138 354 0.354

50 0.0117 297 0.297

60 0.0098 250 0.250

70 0.0083 210 0.210

80 0.0070 177 0.177

100 0.0059 149 0.149

120 0.0049 125 0.125

140 0.0041 105 0.105

170 0.0035 88 0.088

200 0.0029 74 0.074

230 0.0024 63 0.063

270 0.0021 53 0.053

325 0.0017 44 0.044

400 0.0015 37 0.037

There are many variants of sand control methods for open hole, the popular
choices being screens and gravel packs.

11.2 Gravel pack

Gravel packs are also known as external gravel packs, where a sized sand gravel is
placed as an annulus around the sand control screen (Fig. 11-3). Since the 1990s,
open-hole gravel packs became a common form of sand control, particularly in
horizontal wells where gravel packs yield good productivity. The technique is simple;
Chapter 11 Sand Control ·203·

pack the annular space with gravel sized to stop formation sand from being produced,
and size the screen to prevent the gravel from escaping.

Fig. 11-3 Structure of gravel pack

When successfully installed, they prevent the formation from collapsing and
therefore reduce fines production. Generally, in combination with properly packed
gravel around the screen section, open-hole completions offer a higher productivity
index comparing to cased hole systems. While designing gravel pack, Saucier’s rule is
widely adopted and proved successful in field practices. Based on experiments, Saucier
concluded that the gravel size should be 6 times the D50 of formation-particles size
distribution.

11.3 Screens

The techniques applied to control down-hole sand production have not changed for
many years. Almost every sand control solution involves the installation of a
pipe-based filter that is either gravel packed or left as a stand-alone screen. Screen
designs are mainly wire-wrap screen, premium screen, and pre-packed screen.

11.3.1 Wire-wrapped screens


Wire-wrapped screens use spiral-welded corrosion-resistant wires wrapped around a
drilled base pipe to provide a consistent small helical gap, such as 0.012 in (termed 12
gauge) or 0.30mm. A wire wrapped screen is pictured in Fig. 11-4.
·204· Petroleum Production Technology

Fig. 11-4 Wire-wrapped screen

The wire-wrapped screen consists of an outer jacket that is fabricated on special


wrapping machines that resemble a lathe. The shaped wire is simultaneously wrapped
and welded to longitudinal rods to form a single helical slot with any desired width.
The jacket is subsequently placed over and welded at each end to a supporting pipe
base with drilled holes to provide structural support. This is a standard design
manufactured by several companies (Fig. 11-5).

Fig. 11-5 Structure of wire-wrapped screen

Well productivity can be restricted by inflow area. The inflow area depends on the
wire thickness, the slot width and the percentage of screen joint that comprises slots.
For example, in Fig. 11-6 using the (2 times D10) criteria for slot sizing, the screen
inflow areas are calculated for a variety of formation grain sizes and two sizes of wire
Chapter 11 Sand Control ·205·

(0.047 in and 0.09 in). It is assumed that 90% of the screen joint length comprises
slots. Note that if the more conservative D10 criteria is used, then the inflow area
reduces by nearly 50%. Even an inflow area of 5% is more than sufficient if the screens
do not plug. Such an area in an open-hole well is substantially more than the flow area
of a cased and perforated well.

Fig. 11-6 Flow area for wire-wrapped screen

11.3.2 Premium screens


Premium screens are constructed with multiple woven layers (Fig. 11-7 and Fig. 11-8).
An offset layer reduces hot spots both from the outer shroud and through the base pipe.
Premium screens are thinner than pre-packed screens, although the outer shroud makes
them slightly thicker than the wire-wrapped screens. Premium screens typically have
an inflow area of around 30%.

Fig. 11-7 A typical premium screen


·206· Petroleum Production Technology

Fig. 11-8 Internals of premium screen

11.3.3 Pre-packed screens


Pre-packed screens are in fact a modification of wire-wrapped screens. They consist of
a standard screen assembly with a layer of resin-coated gravel, supported by a second
screen (Fig. 11-9). After packing the screen, the complete unit is heated to cure and
harden the resin. The thickness of the gravel layer can be varied to meet special needs.
This screen has a thin lattice screen wrapped around it to prevent gravel from flowing
through the holes in the pipe base before consolidation.

Fig. 11-9 Structure of a pre-packed screen


Chapter 11 Sand Control ·207·

The pre-packed screens offer a good degree of depth filtration, and the relatively
high porosity (over 30%) combined with their very high permeability provide minimal
pressure drops. However, pre-packed screens can be prone to plugging and are no
better at resisting jetting of sand than wire-wrapped screens. Premium screens or
wire-wrapped screens have now largely replaced pre-packed screens, but pre-packed
screens still remain in popular use in some areas of the world.

11.3.4 Expandable screens


Gravel packing has evolved over more than 70 years, but still involves complex fluid
and gravel pumping operations. Stand-alone screens, while being very simple in
operation, do not support the formation, leading to a number of well reliability and
production issues.
Expandable sand screens were developed to overcome the limitations of both
existing techniques, while providing some unique benefits. The first commercial
application of this technology took place in January 1999. Since then, their use has
spread quickly to all parts of the world and in many diverse applications. As of
November 2000, over 23000 ft of expandable sand screen had been run in 25
applications. In 2000, Chevron completed a sidetrack well, using expandable sand
screens for sand control, making it the first case in the Gulf of Mexico.
The development of the expandable sand screen was created by cutting
base-pipe in an overlapping engineered slot pattern around its circumference. The
expandable base pipe is covered longitudinally with overlapping sheets of woven
wire mesh as the filter medium (Fig. 11-10). Tests with these filters showed that
careful selection of aperture size for a given sand mixture effectively minimized
plugging. The filter media is sandwiched between the base pipe and an outer
expandable protective shroud.
Screen expansion is achieved by driving a hardened expansion cone (swage
mandrel) through the base pipe (Fig. 11-11). The base pipe ID expands slightly more
than the OD of the expansion cone. The expansion cone selection is done immediately
prior to running the screen to capture all the available information regarding the
wellbore.
·208· Petroleum Production Technology

Fig. 11-10 Structure of an expandable screen

Fig. 11-11 Expansion of screen by a cone

The slotted tube has several desirable features as the base-pipe for an expandable
screen: ① Upon expansion, its diameter can be increased by over 100%, if necessary,
allowing elimination of large annular gaps. ② It provides a large inflow area when
expanded (typically greater than 30%) that is ideal for a screen containing a highly
porous filter membrane. ③ The expanded base pipe internal diameter is larger than
the swage mandrel diameter. This surplus expansion phenomenon allows the mandrel
to travel freely, but more importantly, allows the expandable screen to achieve a tight
fit inside a confining wellbore, thus providing for wellbore support. ④ An
Chapter 11 Sand Control ·209·

expandable sand screen controls sand production without a frac-pack, but also
overcomes the plugging and erosion problems of stand-alone screens. Moreover, the
elimination of the frac-pack reduces the initial cost as well as the risk of early water
production.
A key component of the system is an expandable sand tight connector, which
allows the entire screen assembly to be expanded to the same dimensions while
allowing flow through it. As the sand screen is expanded, it eliminates the annulus
between the screen and the borehole. This feature effectively permits smaller diameter
sand-face completions for the same production bore size. Besides, reducing the final
bore size can lead to cost reductions of up to 20% on the total well costs. With the
annulus eliminated, it is possible to set devices within the screen to isolate sections or
to squeeze treatment fluids into the formation and there is a dramatic reduction in the
ability of sand particles to move around under producing conditions. In theory, sand
particles are rapidly stabilized, resulting in reduced erosion and screen plugging. These
advantages are depicted in Fig. 11-12.

Fig. 11-12 Advantages of expandable sand screen


·210· Petroleum Production Technology

Attention was required to ensure that the interface between the formation and the
expandable sand screen (perforation tunnels) would maintain an adequate permeability.
Perforation debris in tunnels will compromise their productivity whether the tunnels
are pre-packed or not. The removal of this debris through under-balanced surging was
the initial step to ensure the integrity of the tunnel permeability.
Chapter 12 Flow Assurance
Producing wells often suffer from flow-assurance issues, including scale, wax,
asphaltene and hydrate. These precipitations lead to reduced production, blockage of
wells and pipes, even serious safety hazards. Among these problems, prevention of
hydrate formation and safe removal of hydrate plugs represent 70% of deepwater
flow-assurance challenges, while the remaining 30% deals with scale, waxes, and
asphaltenes. Flow-assurance issues have to be prevented or remediated with chemicals
or well interventions that can be very difficult and costly.

12.1 Scale

Scales are inorganic solids that precipitates from water. They deposit and block
perforations, screens, liners and tubing. Scales also make check valves and safety
valves lose operability. Formation water contains many types of ions at various
concentrations. Typical ions include chloride, sodium, potassium, calcium, magnesium,
strontium, barium, iron, bicarbonate, and sulphate.
These ions were in an equilibrium state after long-term storage in underground
environment. This equilibrium is disturbed by the pressure and temperature changes
induced by production operations. Moreover, invasion of foreign fluids (injected water,
mud, completion fluid, CO2, polymers, surfactants) also breaks this equilibrium and
leads to solids precipitation. Fig. 12-1 shows pipes with scale precipitation.

Fig. 12-1 Scale on pipe wall


·212· Petroleum Production Technology

Common scales are calcium carbonate, barium sulphate, strontium sulphate,


calcium sulphate, and sodium chloride. Carbonate scale can form quickly and block the
pipe within days. The formation of calcium carbonate is affected by pressure,
temperature, water composition and carbon dioxide. Calcite forms as a result of the
reaction of calcium ions with either bicarbonate or carbonate.

Ca 2 + + CO32− → CaCO3 ↓

Ca 2 + + 2HCO3− → CaCO3 ↓ + CO 2 ↑ + H 2 O

In most reservoirs, carbonate ions are rare. Therefore the reaction of calcium and
bicarbonate ions is the major cause of calcite. The reaction can go in either direction
depending on temperature, pressure and pH. For example, if CO2 is removed, the
reaction proceeds and precipitation occurs, as long as the calcium ions are available.
Increase in temperature promotes scale formation. This is similar to while scale
coating a kettle after boiling hard water. Steam flooding is a very popular recovery
method for heavy oil fields. Injection of steam increases the reservoir temperature and
leads to scale formation. When produced fluids flow through an electrical submersible
pump, the temperature increases and scale may form.
Pressure has significant influence on scale formation. Reduction in pressure leads
to less CO2 dissolved in solution and the reaction proceeds to the right hand side. Even
though the mechanism of scale formation is clear, there is little to be done to avoid
scale. The most common approach is to squeeze scale inhibitors into the reservoirs to
prevent scale formation, This aspect will be introduced later.
Calcite scale reacts with acid and dissolves quickly. The most common acid is
hydrochloride acid. Scale problem in tubing can be treated with pumping acid down the
tubing and soaking. Acid treatment also requires corrosion inhibitors. Care should be
taken to minimize potential formation damage by acids and corrosion inhibitors.
Common sulphate scales include barium, strontium and calcium. Mixing
formation water with sea water can result in sulphate scales. Most formation water is
low in sulphate, but sea water is high in sulphate. For example, sea water is injected
into reservoirs for water flooding purpose. Sea water thus mix with formation water in
the well and near wellbore region and scale occurs. Therefore, fresh water instead of
sea water should be injected, or the sulphate ions should be removed from sea water
prior to injection.
Chapter 12 Flow Assurance ·213·

Inhibitor squeeze is the most common method to prevent scale problem. In a


squeeze, scale inhibitor chemical is injected into the formation, typically to an average
10 ft radius from the wellbore. The chemicals first adsorb onto rock matrix, then detach
slowly as normal well production resumes. The squeeze treatment may be preceded by
a preflush to condition the matrix to improve adsorption of the chemicals onto the
matrix. The inhibitor chemical slug is fully displaced into the reservoir often by a light
fluid such as base oil or nitrogen to assist turnaround of the well following squeeze
treatment. It is common practice to allow a soak period following displacement to
optimize adsorption of the inhibitor. This can vary from a few hours to as much as a
day. Common scale inhibitors include inorganic phosphates, organic phosphorous
compounds, polycarboxylic acids, and polyacrylamides.

12.2 Waxes

Waxes or paraffins are long-chain alkanes that solidify at low temperature. Wax
precipitates from the oil phase and deposits on pipe wall. Paraffin deposition reduces
pipe internal diameter, thus leading to reduction in production and even blockage of
pipes. Fig. 12-2 presents a tubing with wax blockage. The wax deposit in the pipe may
look in solid sate, in fact it contains some percentage of oil. When the wax deposit is
young, it contains more than 50% crude oil. While the diffusion process continues, the
oil content decreases and the wax hardens. This is referred to as the aging phenomenon
of wax deposit.

Fig. 12-2 Tubing is blocked by wax

Accurate characterization of wax is possible if the alkane (and other hydrocarbons)


distribution is known. This is determined using a high-temperature gas chromatograph.
It is widely accepted that the paraffin deposition phenomenon is a molecular diffusion
process in nature driven by temperature. When waxy crude flows in a pipe, the
temperature at the pipe wall is the lowest, while the temperature at the pipe center is
the highest. This temperature difference leads to a radial temperature gradient, which
·214· Petroleum Production Technology

produces a concentration gradient of wax in oil. As a result, the concentration of wax is


the lowest close to the pipe wall. The wax concentration gradient serves as the driving
force for wax crystals to diffuse towards pipe wall.
Several parameters are used to define the paraffin behavior. Wax appearance
temperature or cloud point defines the temperature at which wax begins to precipitate.
Because light ends can stabilize paraffin molecules, the cloud point of live oil is
usually lower than that of the dead oil. However, the cloud point of live oil is more
difficult to measure. Thus, cloud point of dead oil samples is often used in thermal
insulation design. This design criterion may be overly conservative. A few methods are
available for cloud point measurement: cold finger, differential scanning calorimeter,
filter plugging, etc. Different methods often produce different results.
Pour point defines the temperature at which the crude oil ceases to flow after
being subjected to a standard rate of cooling. This can present a serious problem when
the offshore well is shut down. The fluid in the well and pipeline cools after the well is
shut down, then wax begins to precipitate and forms a gel. If the fluid temperature falls
below the pour point, it is very difficult to restart a well.
For subsea production system, thermal insulation is the most widely-used method
to mitigate wax. The purpose is to maintain fluid temperature above wax appearance
temperature during steady-state operations. However, when the well is shut down, the
fluid begins to cool and eventually reach sea water temperature after 12 to 36 hours.
The wax deposit during shutdown is not significant, because wax deposition is a slow
process. Thermal insulation not only reduces wax deposits, also prevents hydrates if the
design is valid. Besides, heat tracing can be used on land wells and land pipes to
prevent wax deposition. Pour point depressant and wax inhibitors are polymer or
surfactant-based chemicals that prevent wax deposition to certain extent. They can be
injected through capillary tubes for land wells. For offshore wells, they are often
injected from the tree.
Several strategies are available to remove wax. Hot crude oil can be circulated
down the tubing or the annulus to melt the wax deposits. Scrapers are often used in
land wells to remove wax. Continuous injection of inhibitors is expensive. Therefore,
wax inhibitors are usually limited to startup and shut-down scenarios.
Another popular method is to run a pig through the pipelines to remove wax and
other deposits on pipe wall. There are a few pigs available in the market, such as
simple sphere pigs, foam pigs and smart pigs. The pig is sent into the pipeline from a
Chapter 12 Flow Assurance ·215·

pig launcher (Fig. 12-3, Fig. 12-4) and is pushed by the production fluid or any other
fluids, like dead oil or gas. The pig mechanically scrapes the wax from the pipe wall
(Fig. 12-5).
The pigging operation has to be schedules regularly. If the pigging operation is not
scheduled frequently enough, too much wax may have deposited onto the pipe wall.
During the pigging operation, a pig may get stuck inside the pipe due to the excess
amount of wax in front of it. The pigging frequency is initially determined by lab data
or simulation results. It is later fine tuned based on field practices.

Fig. 12-3 Launching of sphere pigs

Fig. 12-4 Launching of a two-disc pig


·216· Petroleum Production Technology

Fig. 12-5 Pigs return to receiver

12.3 Asphaltenes

Asphaltenes are defined as the heaviest, polar, aromatic compounds of crude oil, with
molecular weights in the range 500 to 1000 or above and specific gravity around 1.3.
An example of asphaltene structure is presented in Fig. 12-6. They are dark deposits

Fig. 12-6 An example of asphaltene molecular structure


Chapter 12 Flow Assurance ·217·

that resemble waxes, however their chemistry is very different from waxes. The polar
nature of asphaltene allows it to dissolve in toluene, but not in normal alkanes. Waxes
melt under high temperature, but asphaltene precipitation is an irreversible process.
Pressure plays an important role in asphaltene precipitation. It is believed that
resins have the ability to stabilize asphaltene particles, preventing them from
aggregation. When the pressure is reduced, the light components expand and resins
migrate to these light-density fractions. After losing the resins as natural stabilizers, the
asphaltenes begin to precipitate and grow. The effect of pressure is depicted in Fig. 12-7.
Therefore, pressure maintenance can prevent asphaltene deposition to some extent.

Fig. 12-7 Effect of pressure on asphaltene precipitation

Apart from pressure and temperature, there are a number of other potential causes
of asphaltene precipitation. The addition of any fluid that reduces liquid density can
promote asphaltene precipitation. The beneficial resins then migrate to the lighter
components and asphaltenes aggregate and precipitate. These scenarios include: ①
mixing fluids form different reservoirs in a well; ② gas lift; ③ miscible gas flooding
with methane or other light hydrocarbon gas; ④ CO2 flooding.
Asphaltene deposition is often independent of asphaltene content. The Clyde field
in the North Sea had serious problems with only 0.5% asphaltene content, whereas
many Venezuelan fields contain up to 5% or even 10% asphaltenes and produce
without any real problems. Asphalt used for road surfaces is a mixture of resins and the
asphaltenes, and is an end product of crude oil distillation. Unstable asphaltenes can
precipitate in the formation, in the tubing or at surface and cause severe restrictions.
Prevention or removal of asphaltene deposits is not easy.
·218· Petroleum Production Technology

Mechanical removal of asphaltene in the tubing is hard work, but is often used in
land or platform wells. In Kuwait, dedicated slick line crews scrape and jar their way
down large numbers of wells in sequence. Hydro-blasting or milling (both with coiled
tubing) will be quicker. Aromatic solvents such as xylene and toluene are highly
effective and will also remove wax deposits.

12.4 Hydrates

Various gases can bond with water molecules and form hydrates under low temperature
and high pressure. Hydrates resemble ice and can block pipes in a relatively short
period of time, as shown in Fig. 12-8 and Fig. 12-9. Hydrates often form in pipelines
when flow is restarted, because of turbulence and mixing of gas and water. Hydrates
also form downstream chokes, due to the Joule-Thompson cooling effect.

Fig. 12-8 Hydrate plug in a pipe

Fig. 12-9 Hydrate forming in a pipe


Chapter 12 Flow Assurance ·219·

Deep ocean temperatures are fairly uniform at about 39°F, except for some
anomalous deepwater current environments. Gulf of Mexico deepwater wellhead
pressures at 15000 psia represent some of the highest in the world. Such a combination
of low temperatures and high pressures provides high driving forces for hydrate
formation.

12.4.1 Types of hydrates


Hydrates are classified by the arrangement of the water molecules in the crystal.
Two types of hydrates are commonly encountered in the petroleum industry: Type I
and type Ⅱ , sometimes referred to as structure Ⅰ and Ⅱ . A third type of hydrate
that also may be encountered is type H (also known as structure H), but it is much
less common.
The simplest of the hydrate structures is the type Ⅰ(Fig. 12-10). It is made
from two types of cages: ① Dodecahedron, a twelve-sided polyhedron where each
face is a regular pentagon, and ② Tetrakaidecahedron, a fourteen-sided polyhedron
with twelve pentagonal faces and two hexagonal faces. The dodecahedral cages are
smaller than the tetrakaidecahedral cages; thus, the dodecahedra are often referred to
as small cages whereas the tetrakaidecahedra cages are referred to as large cages.
Type I hydrates consist of 46 water molecules. If a guest molecule occupies each of
the cages, the theoretical formula for the hydrate is X·5 3/4H2O, where X is the
hydrate former.

Fig. 12-10 Structures of type Ⅰ hydrates


·220· Petroleum Production Technology

Some of the common Type Ⅰ hydrate formers include methane, ethane, carbon
dioxide, and hydrogen sulfide. In the hydrates of methane, carbon dioxide, and
hydrogen sulfide, the guest molecules can occupy both the small and the large cages.
On the other hand, the ethane molecule occupies only the large cages.
The structure of the Type Ⅱ hydrates is much more complicated than that of the
Type Ⅰ. The Type Ⅱ hydrates are also constructed from two types of cages (Fig. 12-11).
The unit structures of a Type Ⅱ hydrate are ① Dodecahedron, a twelve-sided
polyhedron where each face is a regular pentagon, and ② Hexakaidecahedron, a
sixteen-sided polyhedron with twelve pentagonal faces and four hexagonal faces. The
dodecahedral cages are smaller than the hexakaidecahedron cages. The Type Ⅱ hydrate
consists of 136 molecules of water.

Fig. 12-11 Structures of type Ⅱ hydrates

Among the common Type Ⅱ formers are nitrogen, propane, and isobutane. It is
interesting that nitrogen occupies both the large and small cages of the Type Ⅱ hydrate.
On the other hand, propane and isobutane only occupy the large cages.
The pressure-temperature curves for hydrate formation are given in Fig. 12-12 and
Fig. 12-13. on the left side of the curve is the hydrate formation region. In other words,
hydrates begin to form when pressure and temperature fall to the left side of the curve.
The operating pressure and temperature should avoid the hydrate formation region,
otherwise remediation strategies have to be deployed.
Chapter 12 Flow Assurance ·221·

Fig. 12-12 Hydrate formation curves for various gases

Fig. 12-13 Hydrate formation curves for nitrogen and methane


·222· Petroleum Production Technology

12.4.2 Hydrate prevention


Staff should monitor the production system for early hydrate warnings, such as slush in
pigging returns; changes in water rates and fluid compositions at the separator;
pressure-drop increases; and acoustic signals (pinging) of hydrates hitting the
pipelines. On the other hand, pipeline temperature after shutdown should be closely
monitored. After shutdown, fluids in the pipelines cool and hydrate may begin to form.
Common strategies that prevent hydrates include chemical inhibitors, insulation
and heating. Hydrate prevention requires inhibitors that can withstand a substantial
sub-cooling of up to 35°F. Sub-cooling refers to the hydrate equilibrium temperature
minus typical deepwater temperature of 40°F. High-pressure pipelines have high
hydrate-equilibrium temperatures.
Thermodynamic inhibitors are routinely used to shift the hydrate curves. These
chemicals do not affect the nucleation and growth of hydrate crystals. They only
change the hydrate formation pressure or temperature. Two kinds of inhibitors are
commonly used: Methanol and monoethylene glycol (MEG). Effect of methanol on
H2S hydrate curve is depicted in Fig. 12-14. For system pressure of 10 MPa, hydrates
form at temperature below 30℃ in fresh water. By adding 35wt% methanol to fresh
water, hydrate formation temperature is depressed to 12℃. With 50wt% methanol,
hydrate temperature is further depressed to 6℃.

Fig. 12-14 Effect of methanol on hydrate formation curve


Chapter 12 Flow Assurance ·223·

Methanol is flammable and toxic. Moreover, methanol can cause problems for
catalysts in downstream processing. The required dosage of methanol is often high,
resulting in high costs. MEG is also flammable and toxic, and more expensive than
methanol. MEG is easy to recover in the production system, therefore it is routinely
used in wet gas pipelines. A relatively simple and widely-used method to predict effect
of inhibitors on hydrate-forming temperature is the Hammerschmidt equation:

kHW
ΔT = (12-1)
M (100 − W )

where,
T = temperature depression, ℃; kH = constant with value of 1297; W = concentration
of inhibitor in aqueous phase, wt%; M = molar mass of the inhibitor, g/mol.
The Hammerschmidt equation[Eq. (12-1)] predicts the deviation of temperature
from the original temperature without inhibitor, not the hydrate forming conditions.
The Hammerschmidt equation is limited to concentrations of about 30 wt% for
methanol and ethylene glycol, and only to about 20 wt% for other glycols.
Certain low dosage hydrate inhibitors are also popular in the market. Two types
are available: Kinetic inhibitor and anti-agglomerate. Kinetic inhibitors slow down the
hydrate nucleation process and delay the crystal growth process for an extended period
of time, but can not completely prevent hydrates. After the time has lapsed, hydrates
can form very quickly. On the other hand, it is important to make sure that the
inhibitors can function at the system temperature. Anti-agglomerates are polymers and
surfactants that prevent hydrates crystals from accumulating into blockage so that the
flow is not interrupted. They do not delay the nucleation process, but prevent the
growth of crystals. They can be used for sub-cooling above 40ºF. These dispersants are
particularly effective in hydrate-plug protection upon line shut-ins and restarts.
However, even with the low inhibitor concentrations, for long tiebacks between
wellheads and platforms, chemicals typically represent the most expensive
flow-assurance solution.
Hydrate inhibition functions in the aqueous liquid, rather than in the bulk vapor or
oil/condensate. While most of the methanol dissolves in the water, a significant amount
of methanol either remains with the vapor or dissolves into the liquid hydrocarbon
phase. Even though the concentration of methanol in the vapor or liquid hydrocarbon is
small, with low water amounts, the majority of methanol may be consumed by the
·224· Petroleum Production Technology

vapor or liquid hydrocarbons because the hydrocarbon-phase fractions are much larger
than the water-phase fraction.
Insulation can be installed to maintain fluid temperature above the hydrate
forming temperature. One option is to directly cast insulation materials onto the outer
surface of pipes. Another method is to install a concentric pipe outside the oil-gas
pipeline (pipe-in-pipe). The annulus space between the two pipes is filled with
insulation materials. Pipe-in-pipe provides a better insulation, but is more expensive. If
a bare pipe is considered as the baseline cost, insulation can easily double to quadruple
the installed-pipe cost.
In reality, it is estimated that insulation is less efficient than expected
approximately 50% of the time. For example, a recent Gulf of Mexico flow-line
experienced an overall-heat-transfer coefficient of 2 Btu①/(hr ⋅ ft2 ⋅ F), while the design
coefficient was 0.176 Btu/(h ⋅ ft2 ⋅ F), the latter being a typical value for such
applications. Insulation methods are not very effective for gas pipelines because of low
heat capacity of gas.
In combination with insulation, line heating may be done through resistance or
induction heating, with practice favoring the former. Heating costs are very high,
second only to chemical treatment, and the power for heating is generated on platforms,
where typically only 5 to 10 MW may be available. There is evidence that short lines
(less than 20 miles ② ) can be handled with heat management; however, heating
solutions may not be practical at line lengths greater than 50 miles.
After a long shutdown, the hydrate risk is high if the pipeline is restarted with cold
fluids inside. To reduce the risk, hot oil is first circulated in the pipeline to displace the
cold fluids with looped pipes in place. Hot oil also warms up the pipes. This operation
usually takes 5-10 h to complete.

12.4.3 Removal of hydrate plug


Removal of hydrate plug is difficult and time-consuming. Therefore, hydrate issue
should be avoided rather than remediated. The most common way to remove a hydrate
plug from a flow channel is by depressurization. Flow is stopped, and the line is slowly
depressurized from both ends of the plug. At atmospheric pressure, the hydrate
formation temperature is invariably less than that of the surroundings, so heat flows

————————
① 1Btu=1.05506×103J.
② 1mile=1.609344km.
Chapter 12 Flow Assurance ·225·

from the environment into the hydrate plug. The plug melts inward, detaching first at
the pipe wall. This process can takes weeks or even months to complete.
For safety reasons, pressure should be reduced from both sides of the plug. Any
pressure gradient across the detached plug causes it to act like a projectile, with
measured plug velocities up to 180 mi/h for short distances. The hydrate has the
density of ice, almost twice that of the surrounding fluid, so at the line velocity, the
plug momentum is twice that of the surrounding fluids. When the hydrate projectile
encounters an obstruction or change in flow direction, such as a pipe elbow, bend, or
valve, the resulting impact or pressure increase frequently causes line rupture,
equipment damage, fire, and potential injury or loss of life. Hydrate-plug dissociation
should always be done slowly and with great care. Always assume multiple hydrate
plugs; there may be pressure between the plugs.
Coiled tubing may be used to remove a substantial liquid hydrostatic head at the
hydrate face to enable depressurization. Coiled tubing may also be used to inject
methanol or glycol at the face of a hydrate plug, when density is insufficient to drive
the inhibitor to the plug face. But the maximum coiled-tubing distance is currently
approximately 5 miles.
In practice, it may be impossible to depressurize from both sides, when only one
plug end is accessible or when a very long time is required to depressurize a large
upstream volume. In such cases, very careful single-sided dissociation may be done by
experienced personnel.
While heating a plug is not normally an option for a buried or submerged pipeline,
heating should always be done with great care from the ends of the plug. Heating
should be done only with assurance that the plug ends will not contain the pressure.
When hydrates are heated, large confined pressure increases cause pipe ruptures. In
February 2000, an Siberian pipe fitter attempted to remove a hydrate plug by heating
an exposed pipeline with a torch. The pipeline exploded. One man died and four others
were badly injured in the resulting fire.
Chapter 13 Separation and Treatment
The produced fluids normally contain water, oil, gas, plus some solids. After the fluids
reach surface, they need to be separated and treated as soon as possible. Afterwards, oil,
gas and water will receive further treatment and refining. The complete separation and
treatment process is presented in Fig. 13-1.

Fig. 13-1 Separation and treatment of produced fluids

13.1 Separator configurations

All separators abide by similar principles. ① A separator is equipped with an inlet


diverter to achieve initial segregation of gas and liquid. ② A separator must have
adequate length or height to allow liquid droplets to settle from gas stream. At the mean
time, oil droplets coalesce to form an stable oil pad. ③ A separator is equipped with a
mist extractor to capture the small liquid droplets that failed to settle due to gravity.
④ A separator is equipped with necessary controls and valves.
A three-phase separator is required to separate the produced fluids into gas, oil and
water. Three types of separators are available: Horizontal separators, vertical separators
and spherical separators. Horizontal separators are the most popular separators in the
market due to the low cost and high efficiency. The schematic of a horizontal
three-phase separator is presented in Fig. 13-2.
Chapter 13 Separation and Treatment ·227·

Fig. 13-2 A horizontal three-phase separator

The produced fluid enters the separator through the inlet and hits the inlet
diverter. The mixture thus reduces its velocity and achieves initial separation.
Centrifugal inlet device is also common. It makes the incoming stream spin around.
Depending on the mixture flow rate, the reaction force from the separator wall can
generate a centripetal acceleration of up to 500 times the gravitational acceleration.
This action forces the liquid droplets together, then they fall to the bottom of the
separator.
The diverter often contains a down-comer which routes the mixture through the
water phase at the bottom of the separator. This process, named water washing,
promotes water coalescence and reduces turbulence at gas-liquid interface. The
down-comer also helps to reduce the turbulences at the phase interfaces.
The horizontal section allows the oil droplets enough time to rise to the surface
and form an oil pad, while the water phase accumulates at the bottom. The oil pad
flows over the weir and accumulates in the oil chamber. Level controls are required to
maintain oil and water levels. Water and oil have similar densities. They often form an
emulsion (Fig. 13-3) which takes more time to separate.
·228· Petroleum Production Technology

Fig. 13-3 Oil, water and emulsion

Gas is much lighter than oil and water, therefore it is very easy to be separated
from liquid. Gas then flows through the mist extractor and the pressure-control valve.
A wire-mesh type mist extractor is presented in Fig. 13-4. Wire-mesh type mist
extractors are dominant in the market. A vane type mist extractor is presented in Fig.
13-5. The tiny droplets in the gas stream collide with the wires or the vanes and fall
downwards.

Fig. 13-4 Wire-mesh mist extractor


Chapter 13 Separation and Treatment ·229·

Fig. 13-5 Vane mist extractor

A vertical separator is presented in Fig. 13-6. It also employs inlet diverter,


down-comer, mist extractor and necessary level controls. These components serve the
same functions as for the horizontal separator. Vertical separators are often used to treat
well streams with low to intermediate gas–oil ratio and relatively large slugs of liquid.
They handle greater slugs of liquid without carryover to the gas outlet. Vertical
separators occupy less floor space, which is important for facility sites such as those on
offshore platforms where space is limited.
For a single-stage separator, there is an optimum pressure which yields the
maximum amount of oil and minimizes the carry over of heavy components into the
gas phase (a phenomenon called stripping). The yield of oil can be increased by
adding additional separators to the process train, but with each additional separator the
incremental oil yield will decrease (Fig. 13-7).
·230· Petroleum Production Technology

Fig. 13-6 A vertical three-phase separator

Fig. 13-7 Effect of separator stages on oil yield

Capital and operating costs will increase as more separator stages are added to the
process train, so a balance has to be struck between increased oil yield and cost. It is
Chapter 13 Separation and Treatment ·231·

uncommon to find that economics support more than three stages of separation and
one- or two-stage separation is more typical. The increased risk of separation shutdown
is also a contributing factor in limiting numbers. Multistage separation may also be
constrained by low wellhead pressures. The separation process involves a pressure drop,
therefore the lower the wellhead pressure the less scope there is for separation.

13.2 Theories of separation

This section introduces the fundamental theories behind horizontal separator design.
Basically, a separator should allow the liquid droplets in the gas stream to settle. At the
mean time, the oil, water and emulsion are given adequate time to break into two clear
phases of oil and water. Horizontal separators are often designed to be half full of
liquid to maximize the liquid-gas interface. Design of vertical separators is not
included here.

13.2.1 Terminal velocity


After a liquid droplet enters the separator, it falls downwards through the gas stream
and reaches terminal velocity when the gravitational force and drag force acting on it
reach a balance. The droplet terminal velocity is expressed as:
In field units:

ρl − ρg d m
vt = 0.0119 (13-1)
ρg Cd

In SI units:

ρl − ρg d m
vt = 0.0036 (13-2)
ρg Cd

The drag coefficient can be obtained with:


24 3
Cd = + + 0.34 (13-3)
Re Re

where,
ρl = density of liquid, kg/m3;
ρg = density of gas under separator conditions, kg/m3;
·232· Petroleum Production Technology

dm = liquid droplet diameter, μm;


Cd = drag coefficient.
To calculate the terminal velocity, a droplet size must be selected. Based on field
experience, the mist extractor will not be flooded if the droplets larger than 140
microns settle into the liquid phase. Thus the mist extractor can catch the droplets
smaller than 140 μm. The subsequent design is based on the droplet size of 140 μm, but
this criterion can be modified if necessary.
Some separators are designed to remove only very small quantities of liquid that
condense due to temperature or pressure changes in a gas stream that has already
passed through a separator and a mist extractor. These separators, commonly called gas
scrubbers, could be designed for removal of droplets on the order of 500 μm without
flooding the mist extractors.
The separator design must allow enough time for the large liquid droplets to reach
liquid interface. On the other hand, the design also provides enough retention time for
the water droplets to coalesce. It is common to use retention times ranging from 3 to 30
minutes depending upon laboratory or field data. If this information is not available,
the guidelines presented in Tab. 13-1 can be used. Generally, the retention time should
be increased as the oil gravity or viscosity increases.
Tab. 13-1 Recommended oil retention time
Type of crude oil Oil retention time/min
Condensate 2~5

Light oil (30 to 40º API) 5~7.5

Intermediate oil (20 to 30º API) 7.5~10

Heavy oil (less than 20º API) More than 10

Similarly, adequate separator length is required to assure that most of the large oil
droplets in water have sufficient time to coalesce and rise to the oil–water interface.
For the water phase, it is common to use a retention time ranging from 3 to 30 minutes
depending upon laboratory or field data. If this information is not available, a water
retention time of 10 minutes is recommended for the design.

13.2.2 Gas capacity constraint


Now we design a horizontal separator that is half full of liquid. By setting the gas
retention time equal to the time required for a drop to settle to the liquid interface, the
following equations may be derived.
Chapter 13 Separation and Treatment ·233·

In field units:

420TZQg ρg Cd
dLeff = (13-4)
P ρl − ρg d m

where,
d = separator inside diameter, in;
Leff = separator effective length, ft;
P = separator pressure, psia;
T = separator temperature, ºR;
Z = gas compressibility factor under separator conditions;
Qg = gas flow rate, MMscf/d (million standard cubic feet per day).
In SI units:

34.5TZQg ρg Cd
dLeff = (13-5)
P ρl − ρg d m

where,
d = separator inside diameter, mm;
Leff = separator effective length, m;
P = separator pressure, kPa;
T = separator temperature, ºK;
Qg = gas flow rate, m3/h.

13.2.3 Retention time requirement


Adequate separator length is required to assure that most of the large oil droplets in
water have sufficient time to coalesce and rise to the oil–water interface. For the water
phase, it is common to use a retention time ranging from 3 to 30 minutes depending
upon laboratory or field data. If this information is not available, a water retention time
of 10 minutes is recommended for the design.
In field units:

d 2 Leff = 1.42(Qw trw + Qo tro ) (13-6)


·234· Petroleum Production Technology

where,
Qw = water flow rate, stb/d;
trw = water retention time, min;
Qo = oil flow rate, stb/d;
tro = oil retention time, min.
In SI units:

d 2 Leff = 4.2 × 104 (Qw trw + Qo tro ) (13-7)

where,
Qw = water flow rate, m3/h;
trw = water retention time, min;
Qo = oil flow rate, m3/h;
tro = oil retention time, min.

13.2.4 Oil pad thickness


While water droplets settle through oil pad, they soon reach terminal velocity. Based on
the velocity and the specified oil phase retention time, the distance that a water droplet
can settle may be determined. This settling distance establishes a maximum oil pad
thickness given by the following formula.
In field units:

0.00128tro ΔSGd 2m
ho = (13-8)
μ

In SI units:
2
0.033tro ΔSGd m
ho = (13-9)
μ

where,
ΔSG = difference of specific gravity between two phase.
The relationship between oil pad ratio and water phase fraction is shown in Fig.
13-8. The oil pad ratio β is defined as the ratio of oil pad thickness to the separator
diameter. The water phase fraction is the ratio of the area occupied by water phase to
the separator cross sectional area.
Chapter 13 Separation and Treatment ·235·

Fig. 13-8 Oil pad ratio versus water phase fraction

The water phase fraction can be calculated with the formula below.
Aw 0.5Qw trw
= (13-10)
A tro Qo + trw Qw

13.2.5 Water phase thickness


At the mean time, oil droplets rise through water phase to reach oil pad. Based on the
terminal velocity and the retention time. the water phase thickness can be obtained.
In field Units:
51.2trw ΔSG
hw = (13-11)
μw

In SI Units:
1520trw ΔSG
(hw ) max = (13-12)
μw
·236· Petroleum Production Technology

The maximum separator diameter can be calculated with water phase thickness
and the oil pad ratio β.
hw
d max = (13-13)
β

13.2.6 Slenderness ratio


The slenderness ratio for a separator is defined as the ratio of the seam-to-seam length
divided by its outside diameter. Experience indicates that slenderness ratio should range
between 3 and 5 to minimize turbulence. Slenderness ratios outside the 3 to 5 range may
be used but are not as common. Slenderness ratios outside the 3 to 5 range may be used,
but the design should be checked to assure that re-entrainment will not occur.

13.3 Separator sizing

We will design the size of a horizontal separator with the fundamental theories
presented in the previous section.
Example: design the size a horizontal separator. Oil rate = 30 m3/h (standard
conditions); Gas rate = 6000 m3/h (standard conditions); Water rate = 20 m3/h; Oil
specific gravity = 0.85; Gas specific gravity = 0.6; Water specific gravity = 1.05; Oil
viscosity = 10 cP; Water viscosity = 1 cP; Oil and water retention time = 10 min;
Separator pressure = 700 kPa; Separator temperature = 50℃; Drag coefficient = 2;
The difference in oil-water specific gravity:

ΔSG = 1.05–0.85 = 0.2

Oil pad thickness (assume water droplet size as 500 micron):

ho = 0.033×10×0.2×5002/10 = 1650 (mm)

Water phase fraction:

Aw/A = 0.5×20×10/(10×30+10×20) = 0.2

According to Fig. 13-8, β = 0.245


Maximum separator diameter:

dmax = 1650/0.245 = 6735 (mm)


Chapter 13 Separation and Treatment ·237·

Recall the gas capacity constraint (assume droplet size as 100 μm),

dLeff = 6074

Based on this calculation, the separator length is less than 1 m, which is not valid
obviously.
Recall the retention time constraint,

d2Leff = 2.1×107

For the diameter of 72 in (or 1.829 m), the effective length is calculated to be 6.3
m. We need to multiply the effective length by a factor of 1.33 to accommodate the
inlet diverter and mist extractor. Therefore the seam-to-seam length is regarded as 8.4
meters long, and the slenderness ratio equals 4.6. This slenderness ratio is between 3
and 5. Our final design is 1.829 m in diameter and 8.4 m in length.

13.4 Gas treatment

After initial separation, the gas stream still contains water vapor. The gas will go
through dehydration to remove water content, then compression to a higher pressure
for further downstream processing.

13.4.1 Gas dehydration


The gas stream downstream separator still contain water vapor. The water vapor
condenses under low-temperature and high pressure conditions. The water phase in gas
transmission system can cause certain problems. ①Liquid water and natural gas can
form hydrates under low temperature and high pressure. Hydrate may plug pipes and
cause damage to pipes and equipment. ②Presence of water leads to high pressure loss
in pipes and reduces gas transmission efficiency. ③When water is present, CO2 and
H2S become corrosive to pipes.
Solubility of water in natural gas increases with temperature and decreases with
pressure, as presented in Fig. 13-9. Untreated gas often contains several hundred
pounds of water per MMscf (million standard cubic feet) of gas, while gas pipelines
require gas content to be 6~8 lb/MMscf or even lower for offshore pipes. Therefore,
dehydration process is necessary before gas is transported by pipeline.
·238· Petroleum Production Technology

Fig. 13-9 Solubility of water in gas

The most common dehydration methods include adsorption and absorption. In


adsorption dehydration, the water vapor from the gas is concentrated and held at the
surface of the solid desiccant by capillary forces. Solid desiccants have very large
surface areas per unit weight. The most common solid adsorbents are silica, alumina,
and certain silicates known as molecular sieves. Dehydration plants can remove
practically all water from natural gas using solid desiccants. Because of their great
drying ability, solid desiccants are employed where higher efficiencies are required.
A typical solid dehydration plant employs two adsorbing units. While one unit is
adsorbing water, the other unit is being regenerated by hot and dry gas. Under normal
operating conditions, the life of a desiccant ranges from 1 to 4 years. Solid desiccants
become less effective in normal use because of loss of effective surface area as they
age. Abnormally fast degradation is a result of blockage of the small pores and
capillary openings due to lubricating oils, amines, glycols, corrosion inhibitors, and
other contaminants, which can not be removed during the regeneration cycle.
Another dehydration process is absorption by contact with a liquid desiccant such
as glycol. A contact tower is depicted in Fig. 13-10. Dehydration by absorption with
glycol is usually more attractive economically than dehydration by solid desiccant.
Glycols used for dehydration include ethylene glycol (EG), diethylene glycol (DEG),
triethylene glycol (TEG), and tetraethylene glycol (T4EG). TEG has gained nearly
universal acceptance as the most cost-effective glycol for sweet and sour gases over a
wide range of operating conditions.
Chapter 13 Separation and Treatment ·239·

Fig. 13-10 Glycol contact tower for gas dehydration

13.4.2 Gas compression


The compressors used in today’s natural gas production industry fall into two distinct
types: Reciprocating and rotary compressors. Reciprocating compressors are most
widely used in the natural gas industry. They are built for practically all pressures and
volumetric capacities. The cylinder assembly of a reciprocation compressor consists of
a piston, cylinder, cylinder heads, suction and discharge valves, and other parts
necessary to convert rotary motion to reciprocation motion.
A reciprocating compressor is designed for a certain range of compression ratios
through the selection of proper piston displacement and clearance volume within the
cylinder. A typical reciprocating compressor can deliver a volumetric gas flow rate up
to 30000 ft3/min at a discharge pressure up to 10000 psig.
Rotary compressors are divided into two classes: The centrifugal compressor and
the rotary blower. A centrifugal compressor consists of a housing with flow passages, a
rotating shaft on which the impeller is mounted, bearings, and seals to prevent gas from
escaping along the shaft. Centrifugal compressors have few moving parts because only
the impeller and shaft rotate. Thus, its efficiency is high and lubrication oil
consumption and maintenance costs are low. Cooling water is normally unnecessary
·240· Petroleum Production Technology

because of lower compression ratio and lower friction loss. Compression rates of
centrifugal compressors are lower because of the absence of positive displacement.
Centrifugal compressors compress gas using centrifugal force. In this type of
compressor, work is done on the gas by an impeller. Gas is then discharged at a high
velocity into a diffuser where the velocity is reduced and its kinetic energy is converted
to static pressure. Unlike reciprocating compressors, all this is done without
confinement and physical squeezing. In this way, each compressor is required to
develop only part of the station compression ratio. Typically, the volume is more than
100000 ft3/min and the discharge pressure is up to 100 psig.
When selecting a compressor, the pressure–volume characteristics and the type of
driver must be considered. Small rotary compressors (vane or impeller type) are
generally driven by electric motors. Large-volume positive compressors operate at
lower speeds and are usually driven by steam or gas engines. They may be driven
through reduction gearing by steam turbines or an electric motor. Reciprocation
compressors driven by steam turbines or electric motors are most widely used in the
natural gas industry as the conventional high-speed compression machine. Selection of
compressors requires considerations of volumetric gas deliverability, pressure,
compression ratio, and horsepower.

13.4.3 Downstream gas processing


Before gas is delivered to the customers, further processing is normally carried out at
dedicated gas processing plants, which may receive gas from many different gas and
oil fields. Gas piped to such plants is normally treated to prevent liquid dropout under
pipeline conditions. This process is named dew point control. However, the gas may
still contain considerable volumes of natural gas liquids (NGL) and contaminants.
The composition of natural gas varies considerably from lean non-associated gas
which is predominantly methane, to rich associated gas containing a significant proportion
of NGL. NGL refers to the C2~C5+ components remaining once methane and all
non-hydrocarbon components have been removed. Butane (C4H10) and propane (C5H12)
can be further isolated and sold as liquefied petroleum gas (LPG). This is commonly seen
as bottled gas and is a useful method of distributing energy to remote areas.
Sales gas, which is typically made up of methane (CH4) and small amounts of
ethane (C2H6), can be exported by refrigerated tanker rather than by pipeline and has
to be cooled to –160℃ and compressed by a factor of 600. This is then termed
liquefied natural gas (LNG).
Chapter 13 Separation and Treatment ·241·

13.5 Oil dehydration

Produced oil has to be separated from produced water for two main reasons. Firstly,
because the customer is buying oil not water. Secondly, early separation minimizes
costs associated with pumping and corrosion protection, because produced water often
causes corrosion problem for facilities and pipelines. A water content of less than 0.5%
is a typical specification for sales crude. For dehydration of very high-viscosity crudes,
heaters can be used in combination with dehydration tanks. The temperature to which
the crude is heated is a function of the viscosity required for effective separation.
If oil and water are mixed as an emulsion, dehydration becomes much more difficult.
Emulsions can form as oil-in-water or water-in-oil if mixed production streams are
subjected to severe turbulence, as might occur in front of perforations in the borehole or
in a pump. Emulsions can be encouraged to break using chemicals, heat or just gentle
agitation. Chemical breaker is the most common method and laboratory tests would
normally be conducted to determine the most suitable combination of chemicals.

13.6 Water treatment

Water separated from oil usually contains small amounts of oil which have to be
removed before the water can be released to the environment. Industry standards
ranging from 10 to 100 mg/L of oil in water before disposal are currently common. In
most areas, 40 mg/L of oil in water is the legal requirement.
The simplest way to de-oil an oil–water mixture is to use skimming tanks. Over
time the relative density differences will separate the two liquids. On land in situations
where weight and space considerations are not an issue, this can often be accomplished
by using combined settling/storage tanks in a tank farm, either within a field or at an
export facility.
Skimming tanks can reduce oil concentrations down to less than 200 mg/L, but are
not suitable for offshore operations because of time and space constraints. Plate coalesce
vessels or hydrocyclones are generally used offshore and maybe used in tandem with
other water-cleaning devices such as gas flotation units. The choice of de-oiling unit is
influenced by throughput, oil content of the feed, space and weight considerations.
To reduce oil content to levels which meet disposal standards, it is often necessary
to employ rather more sophisticated methods. Two such techniques which can reduce
·242· Petroleum Production Technology

oil in water to less than 40 mg/L are gas flotation and hydrocyclone processes. In a gas
flotation unit, air is bubbled through oily water to capture oil particles which then rise
with the bubble to form a scum at the surface of the flotation unit. The scum can be
removed by rotating paddles. Chemicals are often added to destabilize the inlet stream
and enhance performance.
Hydrocyclones (Fig. 13-11) have become common on offshore facilities and rely
on centrifugal force to separate light oil particles from the heavier water phase. As the
inlet stream is centrifuged, the heavier water phase is ‘spun’ to the outside of the
cyclone whilst oil particles move to the centre of the cyclone, coalesce and are drawn
off upwards. The heavier water is taken out at the bottom.

Fig. 13-11 Hydrocyclone

To ensure disposal water quality is in line with regulatory requirements (usually 40


mg/L), the oil-in-water content is monitored by solvent extraction and infrared
spectroscopy. The specification of 40 mg/L refers to an oil-in-water content typically
averaged over a 1 month period. Oil-in-water standards are generally tightening and whilst
40 mg/L remains acceptable in some areas, 10 mg/L or less is becoming more common. In
closed marine environments such as the Caspian Sea, partially closed environments like the
Arabian Gulf and elsewhere, many companies are re-injecting produced water back into
reservoirs to meet prevailing or self-imposed corporate environmental standards. This
alternative is known as produced water re-injection (PWRI).
Main References

Ahmed T. 2010. Reservoir Engineering Handbook. Oxford: Gulf Professional Publishing


Allen T. and Roberts A. 2012. Production Operations (Fifth Edition). Tulsa: Oil Gas Consultants International and
PetroSkills Publications
Ansari A M. 1994. A comprehensive mechanistic model for upward two-phase flow. SPE roduction & Facilities, 9: 2(2):
143-151
Ansari A. et al., 1994, A comprehensive mechanistic model for two phase upward flow, SPE 28671
Beggs D H, Brill J P. 1973. A study of two-phase flow in inclined pipes. Journal of Petroleum technology, 25(05):
607-617
Beggs D. 2002. Gas Production Operations. Tulsa: Oil Gas Consultants International
Beggs D. 2008. Production Optimization Using Nodal Analysis. Tulsa: Oil Gas Consultants International
Bellarby J. 2009. Well Completion Design. Oxford: Elsevier
Brill J. 1999. Multiphase Flow in Wells. Richardson: Society of Petroleum Engineers
Carroll J. 2003, Natural Gas Hydrates. Oxford: Gulf Professional Publishing
Chilingar G, Yen T. 1983. Some notes on wettability and relative permeability of carbonate reservoir rocks. Energy
Sources, 7(1): 67-75
Colebrook C F, 1939, Turbulent flow in pipes, with particular reference to the transition region between the smooth and
rough pipe laws, Journal of Institute of Civil Engineers, 11(4): 133-156
Dake L.P. 1983. Fundamentals of Reservoir Engineering, Oxford: Elsevier
Duns Jr H, Ros N C J. 1963.Vertical flow of gas and liquid mixtures in wells, 6th World Petroleum Congress, Tokyo
Economides M. and Hill A. 2012. Petroleum Production Systems. New Jersey: Prentice Hall
Fancher G J R, Brown K. 1963. Prediction for pressure gradients for multiphase flow in Tubing, Society of Petroleum
Engineers Journal, 3(1): 59-69
Gao C. 2017. Petroleum Drilling Technology. Beijing: Science Press
Guo B, Ghalambor A. 2012. Natural Gas Engineering Handbook. Houston: Gulf Professional Publishing
Guo B, Liu X. 2017. Petroleum Production Engineering. Oxford: Gulf Professional Publishing
Guo B, Song S. 2013. Offshore Pipelines. Oxford: Gulf Publishing
Hagedorn A., Brown K., 1965, Experimental study of pressure gradients during continuous two phase flow in small
vertical pipe.Journal of Petroleum Technology,17(04): 475-484
Hasan A, Kabir C ,Sarica C. 2002. Fluid Flow and Heat Transfer in Wellbores. Richardson, Texas: Society of Petroleum
Engineers
Jahn F, Cook M. 1998. Hydrocarbon Exploration and Production, Elsevier
Joshi S. 1991. Horizontal Well Technology, Tulsa: PennWell Publishing
Morrow N R.1976.Capillary pressure for uniformly wetted porous media. Journal of Canadian Petroleum Technology,
15(4): 49-69
·244· Petroleum Production Technology

Orkiszewski, J. 1967. Predicting two-phase pressure drops in vertical pipe. Journal of Petroleum Technology, 19(06):
829-838
Poettman F, Carpenter P. 1952. The multiphase flow of gas, oil, and water through vertical flow strings with application to
the design of gas-lift installations.In Drilling and Production Practices, New York: American Petroleum Institute
Stewart M, Arnold K. 2007. Surface Production Operations (Volume 1). Oxford: Gulf Professional Publishing
Treiber L E, Owens W W. 1972.A laboratory evaluation of wettability of fifty oil-producing reservoirs.Society of
Petroleum Engineers Journal, 12(6): 531-540
Appendix A

You might also like