You are on page 1of 234

Fundamentals

of Reservoir Rock
Properties 2 nd Edition
Nayef Alyafei
Fundamentals
of Reservoir Rock
Properties
Nayef Alyafei
nd
2 Edition
First Edition 2019
Second Edition 2021

Hamad Bin Khalifa University Press


P O Box 5825
Doha, Qatar

www.hbkupress.com
www.qscience.com

Copyright © Nayef Alyafei, 2021


Cover image © Nayef Alyafei

All rights reserved.

This book is distributed under the terms of the Creative Commons Attribution-
NonCommercial 4.0 (CC BY-NC 4.0), which permits any noncommercial use,
distribution, and reproduction in any medium, provided the original author(s) and
source are credited.
The online version of this book can be found at:
www.qscience.com

All figures and images by author

ISBN (PB): 9789927137273

DOI: https://doi.org/10.5339/Fundamentals_of_Reservoir_Rock_
Properties_2ndEdition

This title has a supplementary booklet titled, Reservoir Rock Properties Essentials
Supplement [ DOI: https://doi.org/10.5339/Reservoir_Rock_Properties_Essentials ]

Qatar National Library Cataloging-in-Publication (CIP)

Alyafei, Nayef, author.

Fundamentals of reservoir rock properties / Nayef Alyafei. First English edition. – Doha :
Hamad Bin Khalifa University Press, 2019.

pages ; cm

ISBN 978-992-713-727-3

1. Oil reservoir engineering. 2. Petroleum reserves -- Mechanical properties. 3. Geophysics.


I. Title.

TN870.57 .A42 2019


622.3382– dc 23 201927338009
Special acknowledgment to Qatar National Research Fund, QNRF, project number
NPRP10-0101-170086 for funding the open access fees of this book.
This booklet summarizes the different topics covered in this book in a simple and
easy-to-follow format with a one-page article covering each concept. The booklet
highlights each property with a basic knowledge that intends to convey a general
understanding of the covered topics. The purpose of this booklet is to give a quick
overview of the reservoir rock properties. The booklet can be downloaded from the
following link: DOI: https://doi.org/10.5339/Reservoir_Rock_Properties_Essentials
Conversion of Units
Length: Area:
1 ft = 0.3048 m = 12 in 1 ft2 = 0.092903 m2 = 144 in2
1 m = 3.281 ft = 39.37 in = 100 cm 1 m2 = 10.7649 ft2 = 10000 cm2

Mass: Force:
1 lbm = 0.45359 kg 1 lbf = 4.44822 N = 32.2 lbm.ft/s2
1 kg = 2.2046 lbm = 1000 g 1 N = 0.2248 lbf = 1 kg.m/s2

Interfacial Tension: Permeability:


1 N/m = 1000 mN/m = 1000 dyne/cm 1 D = 1000 mD = 9.869233 x 10-13 m2
Volume:
1 ft3 = 0.02831 m3 = 28.3168 L = 0.178 bbl = 0.178 RB
1 m3 = 35.29 ft3 = 1000 L
Pressure:
1 atm = 101.3 kPa = 1.013 bar = 14.696 lbf/in2 (psia)
1 psia = 6.89 kPa = atm/14.696
1 Pa = 1 N/m2 = 1 kg/m.s2 = 10-5 bar = 1.450 x 10-4 lbf/in2 = 10 dyne/cm2
psia = psig +14.7

Density:
1 g/cc = 1000 kg/m3 = 62.427 lb/ft3 = 8.345 lb/gal = 0.03361 lb/in3

Viscosity:
1 cP = 0.01 poise = 0.01 g/cm.s = 0.001 kg/m.s = 0.001 n.s/m2 = 0.001 Pa.s
= 0.01 dyne.s/cm2 = 6.72 x 10-4 lbm/ft.s = 2.09 x 10-5 lbf.s/ft2

Metric Prefixes:

Prefix Symbol Multiplication Factor

giga G 109

mega M 106

kilo k 103

centi c 10-2

milli m 10-3

micro µ 10-6

nano n 10-9

Oilfield Prefixes:

Prefix Symbol Multiplication Factor

Thousand M 103

Million MM 106

Billion MMM or B 109

Trillion T 1012
Index
Chapter 1 Introduction 11

1.1 What is Petroleum? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.2 Origin of Petroleum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.3 Petroleum System. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.4 What is a Reservoir?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.5 Lithology of Petroleum Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . 13

1.6 What is Petrophysics? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14


1.6.1 Routine Core Analysis (RCAL)
1.6.2 Special Core Analysis (SCAL)

1.7 Why Do We Need to Understand Petrophysics? . . . . . . . . . . . . 15

Chapter 2 Porosity 17

2.1 Classification of Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20


2.1.1 Geological Classification of Porosity
2.1.2 Engineering Classification of Porosity

2.2 Calculation of Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.3 Factors Affecting Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22


2.3.1 Primary Factors
2.3.2 Secondary Factors

2.4 Measuring Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24


2.4.1 Laboratory Measurements
2.4.2 Wireline Logging

2.5 Grain/Matrix Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Chapter 3 Rock Compressibility 43

3.1 Types of Rock Compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


3.1.1 Matrix Compressibility
3.1.2 Bulk Compressibility
3.1.3 Pore Compressibility

3.2 Laboratory Determination of Rock Compressibility . . . . . . . . . . 49

3.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

End of Chapter Questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52


Chapter 4 Permeability 55

4.1 Applications of Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.2 Validity of Darcy’s Law for Single-Phase Permeability . . . . . . . . 56

4.3 Darcy’s Law Under Different Boundary Conditions . . . . . . . . . . 57


4.3.1 Case 1: Linear Solution of Darcy’s Law for
Incompressible Fluid
4.3.2 Case 2: Radial Solution of Darcy’s Law for
Incompressible Fluid

4.4 Laboratory Measurements of Absolute Permeability . . . . . . . . 64


4.4.1 Liquid Permeability
4.4.2 Gas Permeability

4.5 Pressure Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71


4.5.1 Pressure Profile: Liquid Flow
4.5.2 Pressure Profile: Gas Flow

4.6 Flow in Layered Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75


4.6.1 Case 1: Linear Flow in Parallel
4.6.2 Case 2: Linear Flow in Series
4.6.3 Case 3: Radial Flow in Parallel
4.6.4 Case 4: Radial Flow in Series

4.7 Flow in Channels and Fractures . . . . . . . . . . . . . . . . . . . . . . . . . . 82


4.7.1 Flow in Channels
4.7.2 Flow in Fractures
4.7.3 Average Permeability with Channels and Fractures

4.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

Chapter 5 Fluid Saturation 97

5.1 Measuring Fluid Saturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99


5.1.1 Extraction Method: Retort Distillation
5.1.2 Extraction Method: Dean-Stark

5.2 Limitations of Using Extraction Methods


to Evaluate Reservoir’s Saturation. . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.1 Drilling Muds
5.2.2 Fluid Properties

5.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108


Chapter 6 Electrical Properties 113

6.1 Understanding Archie’s Law. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113


6.1.1 Introduction to Ohm's Law
6.1.2 Formation Factor
6.1.3 Resistivity Index
6.1.4 Archie’s Equation

6.2 Factors Affecting Resistivity of Reservoir Rocks. . . . . . . . . . . . . . 125

6.3 Measuring Electrical Properties of Reservoir Rocks. . . . . . . . . . 126


6.3.1 Measuring the Formation Factor
6.3.2 Measuring the Resistivity Index

6.4 Applications of Electrical Properties of Reservoir Rocks . . . . . . 128

6.5 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

Chapter 7 Wettability 135

7.1 Understanding Wettability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135


7.1.1 Surface and Interfacial Tension
7.1.2 Adhesion Tension

7.2 Classification of Wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

7.3 Flow Sequence/cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7.4 Measuring Wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140


7.4.1 Contact Angle
7.4.2 Amott Index

7.5 Applications of Wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

7.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

Chapter 8 Capillary Pressure 149

8.1 Capillary Rise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150


8.1.1 Water/Air System
8.1.2 Water/Oil System

8.2 Capillary Pressure Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154


8.2.1 Drainage
8.2.2 Water Re-saturation

8.3 Laboratory Measurements of Capillary Pressure . . . . . . . . . . . . 161


8.3.1 Porous Plate Technique (PP)
8.3.2 Mercury Injection Capillary Pressure (MICP)
8.3.3 Centrifuge

8.4 Capillary Pressure Conversion and Throat Radius Distribution.164

8.5 Leverett J-Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

8.6 Water Saturation Distribution in a Layered System . . . . . . . . . . 172


8.7 Hydrostatic Pressure and Repeat Formation Tester (RFT). . . . . 173

8.8 Applications of Capillary Pressure . . . . . . . . . . . . . . . . . . . . . . . . 176

8.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

Chapter 9 Relative Permeability 183

9.1 Relative permeability Curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

9.2 Recovery Factor Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188


9.2.1 Displacement Efficiency (ED)
9.2.2 Volumetric Sweep Efficiency (EV)
9.2.3 Recovery Factor (RF)

9.3 Laboratory Measurement of Relative Permeability . . . . . . . . . . 192


9.3.1 Steady State (SS)
9.3.2 Unsteady State (USS)

9.4 Three-Phase Relative Permeability . . . . . . . . . . . . . . . . . . . . . . . . 198

9.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

Chapter 10 Data Integration and Volumetric Estimation of


209
Hydrocarbons
10.1 Estimation of Hydrocarbons in Place. . . . . . . . . . . . . . . . . . . . . 209
10.1.1 Net to Gross
10.1.2 Fluid Properties
10.1.3 Layered Systems
10.1.4 Unit Systems

10.2 Data Integration and Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . 218

10.3 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222

End of Chapter Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223

References 232
10
Chapter 1

Introduction
Prior to the discovery of petroleum, mankind used coal as the main source of
energy to operate their machines. Since the first commercial well drilled in the
United States in 1859, the dependence on petroleum as a source of energy has
increased tremendously. From that point onwards, petroleum has been and
will continue to be the main source of energy for decades ahead due to
its availability, efficiency, and low price. In addition, hydrocarbons are not only
used as fuel for our machines, but also as lubricants and raw materials for many
modern industrial products such as plastics, paints, and rubber.

1.1 What is Petroleum?

Petroleum is a naturally occurring hydrocarbon (composed of hydrogen and


carbon atoms) that can exist as a solid, liquid, or gas. The physical state of the
hydrocarbon is a function of the pressure and temperature to which it is exposed
as well as its structure (chain length/molecular weight). However, most of the
hydrocarbons found within the ground are either liquid or gas, and are referred
to as crude oil and natural gas, respectively.

1.2 Origin of Petroleum

There are two theories for the origin of petroleum. They are the organic and
inorganic theories, as stated in Table 1.1.

Table 1.1: Theories for the origin of petroleum.

Organic Inorganic
(derived from living matter, (not derived from living matter)
usually carbon atoms)

States that petroleum evolved States that petroleum was formed


from the decomposition of through chemical reactions
animals and plants that lived between water, carbon dioxide,
during previous geological times. and several inorganic substances
such as carbonates in the earth.

The organic theory is the commonly accepted theory.

11
1.3 Petroleum System

A petroleum system consists of different geological components needed to


generate and store hydrocarbons. These components are source rock, migration
path, reservoir rock, trap, and seal. Source rock is the rock containing organic
matter in sufficient quantity, and is under suitable conditions for the formation
of hydrocarbons. Migration path is the pathway that the hydrocarbons take to
move away from the source rock to the point where they can find a suitable trap.
The forces driving the movement of hydrocarbons out of the source rock come
from tectonic stresses, which are coupled with capillarity (this topic is explained
further in Chapter 8) and buoyancy (density difference); since hydrocarbons are
lighter than water, they move upward. Reservoir rock is the rock that is able
to store hydrocarbons in its pores. The hydrocarbons will continue migrating
upward until they reach a seal. This is an impermeable layer of rock that blocks
the hydrocarbons from further migration. Finally, a trap is a configuration of
rocks, ensuring that the hydrocarbons are stored in it. Traps can be structural,
stratigraphic, or a combination of both. Figure 1.1 shows the components and
processes in a petroleum system.

Seal

Reservoir

Land Plants Aquatic Plants Structural


and Animals and Animals Trap

Burial

Sand Oil and Gas


Migration

Potential Mud Effective


Source Rock Source Rock

Heat

(a) (b)

Figure 1.1: Schematic showing (a) the process of hydrocarbon formation and (b) the migration of
matured hydrocarbon until it reaches an impermeable seal and attains static equilibrium.

1.4 What is a Reservoir?

In petroleum engineering, a reservoir is the place where the hydrocarbons


reside. Our job as petroleum engineers is to access reservoirs and extract the
hydrocarbons (natural gas and/or crude oil) in an economical and environmentally
safe manner. Reservoirs can be classified into three types: oil, gas, and gas-oil

12
reservoirs, as shown in Figure 1.2. Natural gas, if present in a reservoir, is always
on top because it has the lowest density, while water is always at the bottom
because it has the highest density among the three reservoir fluids (gas, oil, and
water).

Oil Well

Oil Well Gas Well


Impermeable
Rock
Impermeable Impermeable
Rock Rock
Gas

Oil Gas Oil

Water Water Water

(a) (b) (c)

Figure 1.2: Schematic showing typical hydrocarbon distributions in (a) an oil reservoir, (b) a gas
reservoir and (c) a gas–oil reservoir.

1.5 Lithology of Petroleum Reservoirs

Lithology is the study of the general physical characteristics of a rock.


Reservoir rocks can be divided into two lithological types, namely, sandstone
and carbonates. Sandstones are formed from grains that have undergone
sedimentation, compaction, and cementation. Carbonates are principally formed
on carbonate platforms by a combination of biogenic and abiogenic processes.

The major characteristics of both sandstone and carbonate rocks are shown in
Table 1.2.

Table 1.2: Geological comparison between sandstone and carbonate rocks.

Sandstone Carbonate

• Usually composed of silica grains • Two major types are


(mainly quartz and some feldspar). limestone (CaCO3) and dolomite
(CaMg(CO3)2).
• Consolidated (the rock is
combined as one unit) or loosely • Pore space consists of inter- or
consolidated. intragranular porosity as well as
areas of dissolution (vugs) and
• May contain swelling clays (clays fractures.
have negative impact on reservoir
quality).

13
1.6 What is Petrophysics?

Petrophysics is the study of rock properties and rock-fluid properties. These


properties, which we will study extensively in the following chapters, include:
porosity, rock compressibility, single-phase permeability, fluid saturation,
electrical properties of reservoir rocks, wettability, capillary pressure,
and relative permeability. Petrophysics can be divided into core and wireline
petrophysics. In this book, we will mainly cover core petrophysics that requires
conducting laboratory experiments on core samples brought from the reservoir
to the surface. Wireline petrophysics, which involves using logs to determine
properties, will also be briefly covered in this book.

Rock samples are extracted from the reservoir through cuttings or coring, can
be subjected to two categories of laboratory analysis: routine core analysis and
special core analysis.

1.6.1 Routine Core Analysis (RCAL)

Routine core analysis attempts to find the basic properties of the reservoir rock
such as porosity, grain density, permeability, and fluid saturation, as shown
below:

Routine Core Analysis


(RCAL)

Porosity Grain Density Permeability Fluid Saturation

1.6.2 Special Core Analysis (SCAL)

Special core analysis is an extension of RCAL, and attempts to measure data that
is more representative of the reservoir conditions. These measurements include
electrical properties of reservoir rocks, wettability, capillary pressure, and relative
permeability, as shown below:

Special Core Analysis


(SCAL)

Electrical Capillary Relative


Properties Wettability Pressure Permeability

14
1.7 Why Do We Need to Understand Petrophysics?

Petrophysics is a fundamental science for petroleum engineers. Most of the


petroleum engineering topics branch out from petrophysical concepts. An
understanding of petrophysical properties helps us in:

• Estimating the quantity of hydrocarbons present in the reservoirs


(e.g. porosity and fluid saturation).

• Understanding how the hydrocarbons will flow from the reservoir to


the well during production (e.g. permeability, wettability, and relative
permeability).

In this book, we will study each petrophysical property extensively.

15
16
Chapter 2

Porosity
Porosity is the ratio of void volume in a porous medium to the total volume of
that medium. Let us assume that we have an empty 350 ml glass, and we fill the
glass to the brim with water to cover the entire volume. Now consider another
identical glass with four ice cubes in it, with each ice cube having a volume of 50
ml. The total volume of ice in the glass will be 200 ml, given that it is not melting.
If we now want to pour water to the glass, we know that there will be room for
just 150 ml of water, since the rest of the volume is occupied by ice. Hence, the
porosity of the glass with the ice cubes will be 150 ml (pore volume, the volume
of water filling the pore space) divided by 350 ml (total volume) and the resulting
porosity will be 0.43. This scenario is shown in Figure 2.1. Basically, porosity
means storage capacity that can indicate the amount of fluid that the porous
medium can store. Porosity can be calculated using the following equation:

Vp
φ= (2.1)
Vt

where ф is the porosity [dimensionless since we are dividing two volumes], Vp is


the pore volume [cm3], and Vt is the total volume [cm3].

Alternatively, we can subtract the matrix volume (in this case, the ice cubes)
from the total volume and divide it by the total volume to obtain the porosity, as
shown in the following equation:

Vt − Vm
φ= (2.2)
Vt

where Vm is the matrix volume [cm3].

Overall, we can say that:

Vp V t − Vm
φ= = (2.3)
Vt Vt

Vt = Vp + Vm (2.4)

Therefore, if we know any two of the volumes, we can calculate the porosity.

17
350 ml 150 ml
of of
water water

(a) (b)

Figure 2.1: Schematic showing (a) a glass filled with 350 ml of water and (b) a glass filled with
water and four ice cubes. As shown, the volume of water in the glass with ice cubes is less since
a matrix volume is present.

Example 2.1
A core sample has a total volume of 24.5 cm3 and a matrix volume
of 18.9 cm3.

(a) What is the pore volume of this sample?


(b) What is the porosity of this sample?

Solution
(a) Equation 2.4 can be used to find the pore volume:

Vt = Vp + Vm

24.5 = Vp + 18.9
Vp = 5.6 cm3

(b) Equation 2.1 can be used to find the porosity:


φ=
Vp
Vt
=
5.6
24.5
= 0.229 or 22.9%

Reservoir rocks are porous and contain fluids in their pores, as shown in Figure
2.2. Porosity measurement from a core is part of RCAL. When we use the term
"core," we usually refer to a cylindrical rock sample with a width and length of a
few centimeters.

18
0.5 mm

Figure 2.2: Schematic showing the pore spaces in a reservoir rock at a micro-scale from a giant
reservoir field. The blue color in the figure represents the water while the black color represents
the matrix.

In addition, when dealing with rocks, we often refer to the matrix volume as the
grain volume (Vg) and the total volume as the bulk volume (Vb). Note that the
fractional porosity value is often multiplied by 100 to make it a percentage;
however, it should always be a fraction when used in calculations. The
porosity of reservoir rocks usually ranges from 5% to 40%. Table 2.1 shows
typical porosity values for different reservoir rocks. The porosity of rocks within a
reservoir indicates how much oil and/or gas is stored in that reservoir. Therefore,
finding the porosity of the reservoir beforehand is important for engineers
because it helps them estimate how economically viable that reservoir is and
how many resources should be invested in it.

Table 2.1: Typical porosity values in reservoir rocks.

Rock Type Range

Loosely consolidated sands 35–40%

Sandstones 20–35%

Well-cemented sandstones 15–20%

Limestones 5–20%

19
2.1 Classification of Porosity

Porosity has two types of classifications: geological and engineering.

2.1.1 Geological Classification of Porosity

In terms of geological classification, porosity is classified into two subdivisions:


primary and secondary. Primary porosity is the original porosity that develops
during the deposition of the material. Primary porosity can be either intergranular
or intragranular (Figure 2.3). Intergranular porosity is the porosity between
grains, while intragranular porosity is the porosity within the grain itself.
Intergranular porosity forms the majority of the porosity of the rock. Secondary
(induced) porosity is developed after deposition by geological processes which
result in vugs and fractures.

Intergranular

Primary

Intragranular

Figure 2.3: Schematic showing the difference between intergranular and intragranular porosities.

2.1.2 Engineering Classification of Porosity

In terms of engineering classification, porosity can be subdivided into two


categories: total and effective. Total porosity (фt) is the total pore volume of the
rock divided by the bulk volume. On the other hand, effective porosity (фe) is
the interconnected pore volume divided by the bulk volume. Ineffective porosity
is the isolated pore volume divided by the bulk volume. Figure 2.4 shows the
difference between effective and ineffective porosity. Usually in sandstones,
фt = фe as they are relatively homogeneous rocks. Carbonate and dolomite
rocks, on the other hand, usually have фt > фe since carbonates are typically
heterogeneous. As petroleum engineers we are mainly interested in the effective
porosity since hydrocarbons can only flow through connected pores.
20
Effective Connected
Porosity Pore

Total
Porosity

Ineffective Isolated
Porosity Pore

Figure 2.4: Schematic showing the difference between porosity subdivisions: total, effective, and
ineffective.

2.2 Calculation of Porosity

If we consider a cubic packing of spheres (ideal situation) and look at a cube


section as shown in Figure 2.5, the length of the cube is 2r, where r is the radius
of the sphere. Thus, the bulk volume of the cube will be:

Vb = (2r)3 = 8r3 (2.5)

The matrix volume in this case is represented by the volume of the spherical
portions in this cubic segment. We have eight equal portions of one-eighth of a
sphere in this cube, thus:

1 4
Vm = 8( sphere) = 1 sphere = πr3 (2.6)
8 3
The porosity then becomes:

V b − Vm 8r3 − 43 πr3 π
φ= = 3
= 1 − = 0.476 (2.7)
Vb 8r 6

We can conclude that the grain size does not affect the porosity of the rock (as
all the radii in the equation cancel out). In other words, having large spheres
or small spheres will lead to the same porosity as long as they are all of the
same size and have the same packing (Figure 2.6). A value of 0.476 is the highest
achievable porosity, and naturally you will always come across a porosity value
less than this.

21
(a) (b) (c)

Figure 2.5: Schematic showing (a) a cubic packing of spheres from which (b) a subset cube is
selected and then (c) analyzed.

(a) (b)

Figure 2.6: Schematic showing (a) the large particle size and (b) the small particle size. Both have
the same porosity as the particle size does not affect the porosity value.

2.3 Factors Affecting Porosity

Porosity can be affected by either primary or secondary factors.

2.3.1 Primary Factors

Particle Packing

Different packing arrangements lead to different porosities, as shown in Figure


2.7. A cubic packing of matrix (Figure 2.7a) leads to a highest possible porosity of
47.6%, as discussed previously, while rhombohedral packing of spheres (Figure
2.7b) leads to a highest possible porosity of 26.0%, which is lower than the
previous case.

22
(a) (b)

Figure 2.7: Different packing leads to different porosities; (a) cubic packing has 47.6% porosity
and (b) rhombohedral packing has 26.0% porosity.

Sorting

Particles are referred to as “well sorted” when they are all of the same size while
they are poorly sorted when they are of different sizes (Figure 2.8). Well-sorted
particles result in a higher porosity compared to poorly sorted particles.

Well Sorted Poorly Sorted


(a) (b)

Figure 2.8: Schematic showing the effect of sorting on porosity in (a) a well-sorted medium and
(b) a poorly sorted medium. A poorly sorted porous medium tends to have a lower porosity than
a well-sorted medium.

23
2.3.2 Secondary Factors

Cementing materials

The presence of more cementing materials means less porosity as there is less
void space available for the storage of hydrocarbons.

Overburden pressure (compaction)

Overburden pressure will lower the pore volume of the rock, leading to lower
porosity. We will cover more of this topic in Chapter 3.

Vugs, dissolution, and fractures

These are formed after deposition and will increase the porosity of rocks.
Dissolution is when the minerals dissolve over time. Some minerals will dissolve
in water. Vugs are large pores formed by dissolution. Fracture is a break or
separation in a rock formation.

2.4 Measuring Porosity

There are usually two methods of measuring porosity. We either measure it


using laboratory measurements (RCAL) at the centimeter scale or using wireline
logging at the meter scale.

2.4.1 Laboratory Measurements

There are several methods of finding porosity in the laboratory. However, we will
focus on the two most common techniques, which are the fluid displacement
and gas expansion using a gas porosimeter. All the rocks used in the laboratory
core analysis are rocks extracted from a reservoir using coring performed by a
downhole instrument, as shown in Figure 2.9.

Figure 2.9: Schematic showing a downhole instrument used to collect rock samples from the
reservoir (sidewall coring tool).

24
Fluid Displacement

The concept of fluid displacement is based on mass/material balance. In this


technique, we weigh a dry core and measure the dimensions, specifically the
diameter and length of the core. Then, we vacuum saturate the core with water
or brine (salt water), for instance, to make sure that the water has filled all the
pore spaces and no air is trapped in the core (Figure 2.10). The core is then
weighed to find the saturated weight. Subtracting the saturated weight from the
dry weight, we obtain the weight of the water in the pore spaces (Figure 2.11). By
dividing the weight of the water by the density of the water, we obtain the pore
volume:
W s − Wd
Vp = (2.8)
ρ
where Ws is the weight of the core saturated with fluid [g], Wd is the dry weight of
the core [g], and ρ is the density of the fluid [g/cm3]; since the fluid in this case is
water, the density is 1 g/cm3. Note that the weight term used in this book is
analogous to mass, unlike in physics where weight is a force.

Example 2.2
A cylindrical core sample has a length of 5 cm and a diameter of 2 cm. The
dry weight of the sample is 56.5 g, and the weight of the sample saturated
with water is 60.3 g. Given that the density of water is 1 g/cm3, find the
porosity of the sample.

Solution
We find the total volume of the core sample:
 2
2
V t = πr 2
L = π 5 = 15.71 cm3
2

We find the weight of water in the sample:


Ww = Ws − Wd = 60.3 − 56.5 = 3.8 g

We find the pore volume of the sample:


Ww 3.8
Vp = = = 3.8 cm3
ρw 1
Equation 2.1 can be used to find the porosity:


φ= =
Vp
Vt
3.8
15.71
= 0.242 or 24.2%

25
Vacuum
Pump
Air
Bubbles

Figure 2.10: Schematic showing the vacuum saturation of a rock sample.

- =

Wsaturated Wdry Wwater


Figure 2.11: Schematic showing the concept of mass/material balance, indicating that subtracting
the dry weight from the saturated weight gives the weight of the liquid in the system. This weight
can be converted to pore volume by dividing it by the density of the liquid in the system.

From the displacement method, we can also find the bulk volume of irregular
shapes. Let us consider a rock with an irregular shape as shown in Figure 2.12.
In order to measure the volume, we need to coat the surface of the rock with
an insulating material such as paraffin (ρ = 0.9 g/cm3) to prevent the fluid we
are using to enter the pores. However, before that we need to measure the
dry weight of the rock sample and the weight of the core with the paraffin. The
difference between the two weights divided by the density of paraffin is the
volume of the added paraffin. This volume will then be subtracted from the final
volume calculation. After performing all these steps, we can find the volume of
the rock using two methods:

1) We can record the initial volume of water in the graduated cylinder,


and then record the new volume of water after submerging the rock. The
difference between the new volume and the initial volume is the volume of
the rock. However, we also need to subtract the volume of the paraffin to
obtain the actual bulk volume of the rock (Figure 2.12a).

26
2) We can also use the Archimedes’ principle to find the bulk volume of the
rock (Figure 2.12b):

Wdf = Wr − Wa (2.9)

where Wdf is the weight of the displaced fluid [g], Wr is the dry weight of the core
or the weight of the core at the initial conditions before submerging (real weight)
[g], and Wa is the weight of the core after submerging it in the fluid (apparent
weight) [g]. Wa is less than the real weight due to buoyancy forces. Note that we
need to suspend the core in order for buoyancy forces to act on the core. Finally,
in order to measure the bulk volume, we need to use the following equation:

Wdf
Vb = − Vcoat (2.10)
ρ

where ρ is the density of the fluid in which the core is submerged [g/cm3], which
in this case is water, and Vcoat is the volume of coat used, usually paraffin [cm3].
Note that Vcoat can be measured by subtracting the weight of the sample with the
coat from the weight of the sample without the coat and divide the product by
the density of coat used, i.e., paraffin.

Example 2.3
The dry weight of a sample is 330 g, and its weight when saturated with
water is 360 g. The apparent weight of this sample in water is recorded as
225 g. Given that the density of water is 1 g/cm3, find the porosity of the
sample. Assuming the sample is coated with a material of negligible weight.

Solution
Equation 2.8 can be used to find the pore volume of the sample:
W s − Wd 360 − 330
Vp = = = 30 cm3
ρw 1
Equation 2.9 can be used to find the weight of the displaced fluid:


Wdf = Wr − Wa = 330 − 225 = 105 g
Equation 2.10 can be used to find the bulk volume:

Wdf 105

Vb = = = 105 cm3
ρw 1
Equation 2.1 can be used to find the porosity:


φ=
Vp
Vb
=
30
105
= 0.286 or 28.6%
27
(a) (b)
Figure 2.12: Schematic showing the two ways of measuring the bulk volumes of irregularly
shaped volumes: (a) using volume difference and (b) using Archimedes’ principle. The gray liquid
represents the increase in fluid level after submerging the rock sample.

The Gas Expansion Method

The second method used to measure porosity is the method of gas expansion
using a helium porosimeter, which relies on Boyle’s law:

P1 V1 = P2 V2 (2.11)

In this method, we usually use helium as it has a low molecular weight and so
can easily enter the smallest pore spaces, which will lead to the most accurate
results. We use the system shown in Figure 2.13a which consists of two chambers
separated by a valve, with a pressure sensor in the first chamber. Chambers
1 and 2 should be of fixed volumes. To break down the process, we need to
understand the following:

1) We fill chamber 1 with helium and then record the pressure; thus, we
have P1 and V1 as shown in Figure 2.13b.

2) If we open the valve to chamber 2, as shown in Figure 2.13c, then Boyle’s


law becomes:

P1 V1 = P2 (V1 + V2 ) (2.12)

3) If we consider an actual case where we have a rock inside chamber 2


(Figure 2.13d), then Boyle’s law becomes:

P1 V1 = P2 (V1 + V2 − Vm ) (2.13)

28
In this case, helium will access all the chambers and the pore spaces. The only
space helium will not access is the matrix volume as it is not porous; using this
technique, we can calculate the porosity.

We will calculate Vm from the equation above, as V1 and V2 are constants and P1
and P2 will be read from the equipment. After finding Vm and also knowing the
bulk volume of the core, which is easy to measure, we can calculate the pore
volume as Vp = Vb – Vm and the porosity is equal to Vp divided by Vb.

It is important to mention that both the fluid displacement and gas expansion
tests measure the effective porosity as fluids can only access the connected
pores.

Example 2.4
A helium porosimeter is used to find the porosity of a certain core sample.
Both the chambers in the porosimeter have a volume of 100 cm3, and
the sample has a bulk volume of 16.2 cm3. Initially, helium is contained in
chamber 1, the sample is placed in chamber 2 and the valve separating the
two chambers is closed. The initial pressure in chamber 1 is recorded to be
30 kPa, and the pressure after the valve is opened is recorded to be 16 kPa.
Find the porosity of the core sample.

Solution
Equation 2.13 can be used to find the matrix volume:

P1 V1 = P2 (V1 + V2 − Vm )

Rearranging this equation:


P1 V 1
Vm = V1 + V2 −
P2
30 × 100
Vm = 100 + 100 − = 12.5 cm3
16
We find the pore volume:

Vp = Vb − Vm = 16.2 − 12.5 = 3.7 cm


3

The porosity can be found using Equation 2.1:

φ = Vp = 3.7 =
Vt 16.2
0.228 or 22.8%

29
(1) (2)

(a)

(b)

(c)

(d)

Figure 2.13: Schematic showing a helium porosimeter where (a) the system has two chambers
separated by a valve with a pressure sensor on chamber 1, (b) helium is introduced in the first
chamber and the pressure reading is taken; the volumes of the chambers need to be taken
before the actual measurement as part of the calibration process, (c) the valve is opened and
helium reaches the second chamber and, in this case, the pressure will decrease as the initial
helium introduced in the system has a larger volume to occupy (both V1 and V2), and (d) all the
previously discussed steps are incorporated and measurements are taken.

30
2.4.2 Wireline Logging

Wireline logging is the acquisition and analysis of petrophysical properties as a


function of depth. Figure 2.14 shows a schematic of a logging tool (sonde) used
in petrophysical measurements in the well. Wireline logging is usually referred
to as taking in situ measurements inside the well. Different wireline tools are
used to acquire three main properties: lithology, porosity, and fluid saturation.
For lithology, gamma rays from formation rocks as a result of the decay of
radioactive elements are recorded; shales have more radioactive components
than reservoir rocks. For fluid saturation, we use resistivity logs, which are based
on the concept that hydrocarbons have higher resistivity (lower conductivity)
than water. We will cover this in more detail in Chapter 6.

Well Log
SP Resistivity

Increasing Increasing Increasing


Radioactivity Resistivity Porosity

Shale

Oil Sand

Shale
Gamma Ray Resistivity Porosity

Figure 2.14: Schematic showing a wireline logging tool (sonde) and a typical logging response for
three properties.

Porosity Logs

Porosity logs can be divided into three types:

• Bulk Density Log (Density Log).


• Neutron Log.
• Acoustic (Sonic) Log.

31
Generally speaking, several logs are run simultaneously on the same logging
tool for several measurements to be taken at the same time. Porosity logs do
not measure porosity directly; porosity is instead obtained by performing some
calculations on the log data. Moreover, porosity logs measure the total porosity
when compared to laboratory measurements, because logging tools can access
both the isolated and connected pores.

Bulk Density Log (Density log)

A density log records the bulk density of the porous media near the well against
the depth. Moreover, it uses a radioactive source to generate gamma rays. The
gamma rays collide with electrons in the rock formation, losing energy. The
detector in the logging tool measures the intensity of the back-scattered gamma
rays, which is related to electron density of the formation. Electron density is a
measure of the bulk density. The bulk density is dependent on:

• Density of the lithology


• Porosity
• Density and saturation of the fluids in the pores

Again, the log reads bulk density, which is explained further in the following
equation:

ρb = ρm (1 − φ) + ρf φ (2.14)
   
Matrix Fluids in the
pore spaces
where ρb is the bulk density [g/cm3], ρm is the matrix density [g/cm3], which is
constant depending on the formation (the typical values will be discussed later),
ф is the porosity, and ρf is the density [g/cm3] of the fluid that occupies the pore
space and can be either water, oil, or gas; the fluid density is also constant.

Then, we can rearrange the equation to obtain:

ρm − ρb
φ= (2.15)
ρm − ρf

An example of a density log (RHOB) is shown in Figure 2.15.

Neutron Log

The logging tool emits high energy neutrons into the rock formation and the
neutrons collide with the nuclei of the formation’s atoms. The neutrons lose
energy (velocity) with each collision; the most amount of energy is lost during
collisions with a hydrogen atom nucleus. The resulting low energy neutrons are
detected and their count rate is related to the number of hydrogen atoms in
the formation. The higher the count rate, the lower the porosity, and vice versa.
Sometimes corrections have to be made to the readings based on the lithology
32
under study. An example of a neutron porosity log (NPHI) is shown in Figure
2.15.

Acoustic (Sonic) Log

The sonic tool usually consists of sound transmitters and receivers. The concept
relies on sound traveling at different speeds, depending on the medium being
solid, liquid, or gas. Sound travels faster in solids than in liquids.

To calculate the porosity from the sonic log, we use the following equation:

∆TL = ∆Tm (1 − φ) + ∆Tf φ (2.16)


     
Matrix Fluids in the
pore spaces

where ∆TL is the interval transit time read from the log [µs/ft], ∆Tm is the interval
transit time through the matrix [µs/ft], and ∆Tf is the interval transit time through
the fluid [µs/ft]; for water, this time is 190 µs/ft. Interval transit time is the
time taken for the sound wave to travel a certain distance, proportional to the
reciprocal of velocity between the transmitter and the receiver.

Then, we can rearrange the equation to obtain:


∆Tm − ∆TL
φ= (2.17)
∆Tm − ∆Tf

∆Tm is a constant that is based on the lithology (Table 2.2).

Table 2.2: Expected travel time through the matrix for different rocks.

Lithology ΔTm [μs/ft]

Sandstone 55.5

Carbonate 47.5

Dolomite 43.5

An example of a sonic log (DT) is shown in Figure 2.15.

Responses of Porosity Logs

The three porosity logs respond differently to different matrix compositions and
to the presence of gas and oil. The combination of these logs along with the
resistivity and lithology logs is very important to understand the full picture of
the reservoir. The above sections briefly introduce the three logs that are used
to estimate porosity. A detailed interpretation of petrophysical properties from
logs is beyond the scope of this book.

33
2.5 Grain/Matrix Density

Grain/matrix density [g/cm3] is also considered a part of the RCAL. In order to


measure it, we need to know the bulk volume, the weight of the rock sample,
and the pore volume. Finding the matrix density can easily be a part of porosity
measurement as only the weight of the sample will be required. After we measure
the porosity, we can find the matrix volume and hence the matrix density can be
obtained through the following equation:

Wm
ρm = (2.18)
Vm

where ρm is the density of the matrix [g/cm3], Wm is the weight of the matrix or the
dry weight of the core [g] as the density of air is assumed to be negligible; Wm =
Wd (dry weight of the entire core), and Vm is the volume of the matrix [cm3], which
is the bulk volume minus the pore volume (Vb – Vp).

Typical matrix densities of different rock types are shown in Table 2.3.

Table 2.3: Typical matrix densities for different rock types.

Lithology ρm [g/cm3]

Sandstone 2.65

Limestone 2.71

Dolomite 2.87

34
RHOB
Depth, 1.95 g/cc 2.95 DT
ft NPHI 150 μs/ft 50
0.45 -0.15

10700

10800

10900

Figure 2.15: Three different wireline logs: density log (RHOB), neutron log (NPHI), and sonic log
(DT).

Example 2.5
a) Using the density log from Figure 2.15, calculate the porosity of a
sandstone formation at a depth of 10,860 ft. Assume that the formation is
saturated with water having a density ρ = 1 g/cm3.

35
b) Using a neutron log (Figure 2.15), calculate the porosity at a depth of
10,720 ft.

c) Using an acoustic log (Figure 2.15), calculate the porosity of a sandstone


formation at a depth of 10,820 ft. Assume that the formation is saturated
with water having a ∆Tf = 195 µs/ft.

Solution
a) From Figure 2.15, it can be seen that the density log scale has 10
increments. The value of ρb varies from 1.95 to 2.95 g/cm3, thus having a
range of 1.00 g/cm3. To obtain the ρb value after each increment, the range
is divided by the number of increments. Therefore, after every increment,
the value of ρb increases by 0.10 g/cm3 from left to right. At a depth of 10,860
ft, ρb is read to be approximately 2.25 g/cm3.

The value of ρm for sandstone is 2.65 g/cm3 (from Table 2.3) and the value
for ρf is given. Using Equation 2.15, the porosity can be found:
ρm − ρb
φ=
ρm − ρf

φ=
2.65 − 2.25
2.65 − 1
= 0.24 or 24%
b) For the neutron log, the porosity value can be read directly from the log.
From Figure 2.15, it can be seen that the ф value varies from 0.45 to -0.15,
thus having a range of 0.60. Since there are 10 increments, the value of ф
decreases by 0.06 from left to right after every increment. At a depth of
10,720 ft, the value of ф is read to be approximately equal to 0.15 or 15%.

c) From Figure 2.15, it can be seen that the ∆TL value varies from 150 to 50,
thus having a range of 100. Since there are 10 increments, the ∆TL value
decreases by 10 from left to right after every increment. At a depth of 10,820
ft, the ∆TL value is read to be approximately 70 µs/ft.

The value of ∆Tm for sandstone is 55.5 µs/ft (from Table 2.2) and the value
for ∆Tf is given. Based on these:
∆Tm − ∆TL
φ=
∆Tm − ∆Tf

φ=
55.5 − 70
55.5 − 195
= 0.1 10%
or

36
2.6 Summary

Porosity is the ratio of void volume in a porous medium to the total volume of
that medium. It is measured using the following equation:
Vp V b − Vm
φ= = (2.19)
Vb Vb
It is important to measure porosity accurately as it helps in quantifying the
amount of hydrocarbons stored within a reservoir. In terms of geological
classification, porosity is classified as primary and secondary. In terms of
engineering classification, porosity is subdivided into total and effective. Porosity
can be affected by several primary factors, such as particle packing and sorting.
It is also affected by secondary factors such as cementing materials, overburden
pressure, vugs, dissolution, and fractures. Porosity can be measured in the
laboratory using the fluid displacement method and the gas expansion method
using a helium porosimeter. In the field, porosity can be measured using wireline
logging, where measurements are taken inside wells. Three different porosity
logs, namely bulk density log, neutron log, and acoustic log, are used to estimate
the porosity in the formation adjacent to the well. The matrix density for different
types of rocks can also be obtained from the porosity measurements.

Table 2.4 summarizes the concept of porosity.

Table 2.4: Definition of porosity and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Porosity ф The fraction of the Porosity is important


bulk volume of a to quantify the
material (rock) that amount of gas and/
is occupied by the or oil in the reservoir.
pores (voids) in
that material.

37
End of Chapter Questions

Question 2.1

A 200 cm3 beaker is completely filled with sand, and 37 cm3 of water is added to
the beaker until it reaches the top of the sand and no more can be added. What
is the porosity of the sand?

Question 2.2

A core sample has a cubic packing of spheres as shown in Figure 2.5c. Find the
matrix volume of this sample if the bulk volume of this sample is 216 mm3.

Question 2.3

A cylindrical core sample has a length of 8 cm and a diameter of 3 cm. The dry
weight of the sample is 90.8 g, and the weight of the sample saturated with oil of
density 0.75 g/cm3 is 101.2 g. What is the porosity of this sample?

Question 2.4

The fluid displacement technique is used to find the porosity of a certain core
sample. The sample is first coated with paraffin (density of 0.9 g/cm3) and
submerged in a beaker filled with water (density 1 g/cm3), where the volume of
water before and after submerging the sample is recorded. Then, the paraffin is
removed, the sample is saturated with water and then weighed. The following
information is given for the sample:

- Dry weight of the core = 140.2 g


- Weight of the core with paraffin = 145.6 g
- Initial volume in the beaker = 242.5 cm3
- Final volume in the beaker = 318.0 cm3
- Weight of the core saturated with water = 156.8 g

For this sample, find the:

a) Bulk volume

b) Pore volume

c) Porosity

d) Matrix density and the lithology

38
Question 2.5

The dry weight of a sample is 241.0 g, and its weight when saturated with a
particular oil is 266.0 g. The apparent weight of this sample in that oil is recorded
as 143.1 g. Given that the density of this oil is 0.82 g/cm3, find the

a) Porosity of the sample

b) Matrix density and the lithology of the sample

Question 2.6

A core has a porosity of 0.28. The dry weight of the core is 156.4 g, and the weight
of the core when saturated with a 0.75 g/cm3 oil is 175.9 g.

a) What is the pore volume of the core?

b) What is the bulk volume of the core?

c) What would the apparent weight of the dry core be when it is


immersed in the given oil if the core is coated with a material of negligible
weight and volume?

d) When the dry core is coated with paraffin (density 0.9 g/cm3), its
weight in air is recorded as 166.1 g. What would the apparent weight of
the coated core be when immersed in water (density 1 g/cm3)?

Question 2.7

A core sample is placed inside a porosimeter. Both the chambers in the


porosimeter have a volume of 150 cm3, and the sample has a bulk volume of 41.4
cm3. Initially, helium is contained in chamber 1, the sample is placed in chamber
2, and the valve separating the two chambers is closed. After the valve is opened,
a 44% decrease in the pressure reading is recorded compared to when the valve
is closed.

a) What is the porosity of this sample?

b) In the same porosimeter, a different sample with the same bulk volume
is now used with a porosity of 0.15. Given that the initial pressure before
the valve is closed is recorded as 25 kPa, what is the final pressure after
the valve is opened?

39
Question 2.8

For this question, use the logs shown in the figure below. Find the porosity of
the dolomite formation using the density log, neutron log, and acoustic log at a
depth of:

a) 5024 ft

b) 5044 ft

c) 5122 ft

Assume that the formation is saturated with an oil having a density ρ = 0.85 g/
cm3 and a travel time ∆Tf = 325 µs/ft.

Question 2.9

i. The following data is available for core sample A extracted at a depth of 5046 ft
(see the figure shown below):

- Bulk volume Vb = 25.31 cm3


- Grain volume, Vg = 19.19 cm3
- Dry weight of the core sample, Wd = 50.85 g

a) What is the porosity of this core sample?

b) What is the grain density of this sample?

c) Given that the pores of the rock are filled with brine of density
1.20 g/cm3, find the porosity of the rock from the density log.

ii. Core sample B was extracted at a depth of 5099 ft (see the same figure). The
extracted core was coated with paraffin and immersed in a container of liquid.
Upon immersion, 12.30 cm3 of the liquid was displaced. The dry weight of this
sample was measured to be 25.30 g, and the weight of the sample coated with
paraffin was measured to be 27.13 g. The density of paraffin is 0.9 g/cm3.

a) What is the bulk volume of this sample?

b) The paraffin was removed from the core, and the core was
subsequently saturated with oil of density 0.85 g/cm3. The saturated
weight of the core is 26.53 g. What is the porosity of this sample?

c) What is the grain density of this sample?

d) Given that the pores of this rock are filled with brine of interval transit
time 190 µs/ft, find the porosity of this rock from the sonic log.

40
iii. Determine the lithology of core samples A and B, and give two justifications
for each.

RHOB
Depth, 2.0 g/cc 3.0 DT
ft NPHI 140 μs/ft 40
0.3 -0.10

5020

5040

5060

5080

5100

5120

41
42
Chapter 3

Rock
Compressibility
The concept of rock compressibility is similar to that of squeezing a sponge.
Sponges are highly porous media used for many applications, including washing
dishes due to their high capacity to store water. Pressure is applied to the wet
sponge to reduce the pore volume and push the water out. The higher the
pressure applied, the higher the reduction in the pore volume, resulting in more
water coming out. Similarly, rocks are subjected to an overburden (or compaction)
pressure that reduces their pore volume (Figure 3.1). If we analyze this figure, the
pore is subjected to three types of pressure, one external and two internal. The
external pressure is the overburden pressure Pov, which increases with burial
depth and is generally around 1 psig/ft. The internal pressures include the fluid
pressure Pf, which is the pressure exerted by the fluid in the pore spaces (around
0.465 psig/ft for brine), and the matrix pressure Pm, which is the resistance to
deformation of matrix material. For instance, it would be easier to squeeze a
sponge than an iron ball. When we have a system at equilibrium, downward
pressure forces are equal to the upward pressure forces, and we have:

Pov = Pf + Pm (3.1)

where Pov is the overburden pressure [psig], Pf is the fluid pressure [psig], and Pm
is the matrix pressure [psig].

Both Pf and Pm are offsetting part of the overburden pressure. In the reservoir,
as we produce hydrocarbons or water, the fluid pressure decreases as the fluids
leave the reservoir. This reduction in fluid pressure leads to a reduction in pore
volume and porosity, thereby reaching a new equilibrium point. Porosity is a
function of the degree of compaction; the higher the compaction, the lower the
porosity (Figure 3.2). In addition, compaction is a function of burial depth; the
greater the burial depth, the greater the degree of compaction. Compaction
effects are generally irreversible after the rock is compacted; the pore volume
will not reverse significantly even though we reduce the pressure. This is because
compaction pressure is high enough to cause an inelastic deformation (plastic
deformation) of the rock. Inelastic deformation occurs when a high pressure
causes a permanent change in an object’s shape. This is different from elastic
deformation, which is a temporary change in the object’s shape when it is
subjected to lower ranges of pressure. Stretching a rubber band is an example
of an elastic deformation. This phenomenon makes laboratory measurements
generally accurate, as the pore volume does not need to be altered any further

43
after extracting the core from the reservoir conditions. The relative change in
pore volume with changes in compaction pressure in the reservoir is known as
rock compressibility.

Pov

Pm Pf

Figure 3.1: Schematic showing pressure forces acting on a pore. The figure indicates a system
at equilibrium where the overburden pressure (Pov) is equal to the sum of fluid (Pf) and matrix
(Pm) pressures.

(a) (b)
Figure 3.2: Schematic showing the effect of compaction pressure on pore volume, at (a) initial
rock condition and (b) after a high overburden pressure is applied, which leads to lower pore
volume and thus lower porosity.

3.1 Types of Rock Compressibility

Rock compressibility is divided into three major types: matrix, bulk, and pore.
The general form of rock compressibility is:
 
1 dV
c=− (3.2)
V dP T

where c is the rock compressibility [1/psig], V is the volume [ft3], and dV/dP is the
change in volume over change in pressure [ft3/psig]. The subscript T indicates
that the compressibility is calculated at a constant temperature (isothermal),
assuming constant temperature in the reservoir. Moreover, the negative sign
is to make the rock compressibility positive; dV/dP is negative since increasing
pressure reduces the volume.

44
Rock compressibility is usually referred to as isothermal rock
compressibility as petroleum reservoirs can be considered isothermal
(constant temperature).

3.1.1 Matrix Compressibility

Matrix compressibility is the fractional change in matrix volume per unit change
in pressure; its equation is given below:
 
1 dVm
cm =− (3.3)
Vm dP T

where cm is the matrix compressibility [1/psig], Vm is the matrix volume [ft3], and
dVm/dP is the change in matrix volume over change in pressure [ft3/psig].

Since it is difficult for the matrix volume to change with pressure, we usually
ignore the matrix compressibility and assume it to be negligible:

∆Vm ∼
= 0, cm ≈ 0 (3.4)

3.1.2 Bulk Compressibility

Bulk compressibility is the fractional change in bulk volume per unit change in
pressure; its equation is given below:
 
1 dVb
cb = − (3.5)
Vb dP T

where cb is the bulk compressibility [1/psig], Vb is the initial bulk volume [ft3], and
dVb/dP is the change in bulk volume over change in pressure [ft3/psig].

Bulk compressibility is important for ground subsidence studies, especially in


reservoirs close to the surface with rapid production (rapid reduction in fluid
pressure). Subsidence is the gradual caving in or sinking of a piece of land which
has environmental impacts.

3.1.3 Pore Compressibility

Pore compressibility is the fractional change in pore volume per unit change in
pressure. Pore compressibility is also known as formation compressibility.
The two terms can be used interchangeably and their equation is given below:
 
1 dVp
cp/f = − (3.6)
Vp dP T
where cp/f is the pore/formation compressibility [1/psig], Vp is the initial pore
volume [ft3], and dVp/dP is the change in pore volume over change in pressure
[ft3/psig].

45
Example 3.1
A rock sample has a pore volume of 18 cm3. This pore volume decreases by
0.15 cm3 as a pressure difference of 900 psig is applied on the sample. What
is the pore compressibility of this sample?

Solution
Equation 3.6 can be used to find pore compressibility:
 
1 dVp
cp = −
Vp dP T
 
1 −0.15
cp = − = 9.26 × 10−6 psig−1
18 900

We know from Chapter 2 that Vb=Vm+Vp, therefore:

∆Vb = ∆Vm + ∆Vp (3.7)

We also know from Equation 3.4 that the change in matrix volume is negligible,
therefore:

∆Vb ≈ ∆Vp (3.8)

This means that a change in the pore volume will lead to an equivalent change
in the bulk volume.

Pore/formation compressibility is important for reservoir performance


estimation, as a reduction in reservoir fluid pressure will lead to a reduction in
pore volume and thus an increase in the extraction of hydrocarbons. It is worth
mentioning that in hydrocarbon production through natural pressure depletion
(reduction), the pressure difference does not lead to significant changes in pore
volume and thus the effect of compressibility might be ignored in some cases.

Rock compressibility is not part of either RCAL or SCAL; however, it is an important


rock property to estimate hydrocarbon production due to compaction.

46
Example 3.2
The bulk volume of a rock sample is 150 cm3 and its porosity is 18% at
ambient conditions. The pore compressibility of this sample is 7×10-6 psig-1.
Given that the overburden pressure gradient is 0.95 psig/ft, what would be
the porosity of this sample at a depth of 8000 ft?

Solution
We find the overburden pressure at a depth of 8000 ft:

Pov = 8000 × 0.95 = 7600 psig

The change in pressure dP relative to the pressure at the ground level is


7600 psig.

We find the pore volume:


Vp = φVb

Vp = 0.18 × 150 = 27 cm
3

Equation 3.6 can be rearranged to find the change in pore volume:

dVp = −cp Vp dP
dV = −7 × 10−6 × 27 × 7600 = -1.44 cm
3
p

We now find the new pore volume at a depth of 8000 ft:


Vp,new = Vp, old + dVp

Vp,new = 27 − 1.44 = 25.56 cm
3

The bulk volume will also change due to this change in pore volume. Since
the matrix volume remains unchanged, the change in bulk volume is the
same as the change in the pore volume.

We find the new bulk volume at a depth of 8000 ft:


Vb,new = Vb, old + dVp

3
Vb,new = 150 − 1.44 = 148.56 cm

We find the new porosity:


Vp,new
φnew =
Vb,new

25.56
φnew = = 0.172 or 17.2%
148.56

47
Example 3.3
A reservoir of 1300 ft outer radius and 0.35 ft inner radius with 42 ft
thickness experiences a drop in fluid pressure from 5300 psig to 4200 psig.
The initial porosity of the reservoir is 16.4%. What is the new porosity after
the pressure drop given that the formation compressibility is 4×10-5 psig-1?

Solution
We find the bulk volume:

Vb = πr2 L

Vb = π × (13002 − 0.352 ) × 42 = 2.23 × 10
8
ft
3

We find the pore volume:


Vp = φVb

Vp = 0.164 × 2.23 × 108 = 3.657 × 10
7
ft
3

Equation 3.6 can be rearranged to find the change in pore volume:

dVp = −cp Vp dP

dVp = −4 × 10−5 × 3.657 × 107 × 1100 = -1.609 × 10
6
ft
3

We now find the new pore volume after the pressure drop:
Vp,new = Vp,old + dVp

Vp,new = 3.657 × 107 − 1.609 × 106 = 3.496 × 10


7
ft
3

The bulk volume will also change due to this change in pore volume. Since
the matrix volume remains unchanged, the change in bulk volume is the
same as the change in the pore volume.

We find the new bulk volume after the pressure drop:

Vb,new = Vb,old + dVp

Vb,new = 2.23 × 108 − 1.609 × 106 = 2.214 × 10


8
ft
3

We find the new porosity:


Vp,new
φnew =
Vb,new

3.496 × 107
φnew = = 0.158 or 15.8%
2.214 × 108

48
3.2 Laboratory Determination of Rock Compressibility

The experimental set-up for measuring rock compressibility is shown in Figure


3.3. We need to measure the porosity and pore volume of the sample prior to
starting the rock compressibility experiment. Then, we saturate the core with
100% water, for instance, and insert it in a core holder. After that, we gradually
increase the surrounding pressure, which in this case is the oil pressure. For
simplicity, we first start with atmospheric pressure, which does not cause any
change in the pore volume. Then, we increase the surrounding pressure to 1000
psig, which will cause the pore volume to decrease and thus expel water to the
graduated pipette attached to it. The volume expelled represents the change in
pore volume while the difference between the atmospheric pressure and the
1000 psig represents the change in pressure. We need to wait until the system
reaches an equilibrium, and no further water is expelled before we further
increase the pressure to obtain the second data point. After obtaining several
data points, we can plot the cumulative change in pore volume over the initial
pore volume, fractional volume [∆Vcp/Vp], as a function of the applied pressure.
The slope of the linear portion of the graph is the pore compressibility, as shown
in Figure 3.4.

10
Graduate
Pipette 20

30

40

Pressure 50

Gauge
60

70

80

90
0 00
15

0 00
30
100

Oil Water
Core

Pump

Figure 3.3: Schematic showing the experimental set-up for measuring rock compressibility.

49
∆Vcp Slope = cp
Vp

P [psia or psig]

Figure 3.4: A plot representing fractional volume as a function of applied pressure. The slope of
this curve is the pore compressibility. We can see that after a certain point, the slope becomes
constant and can be considered as the average pore compressibility.

Example 3.4
The following table shows the results when pressure is applied on a rock
sample with fluid in its pores:

Pressure applied Cumulative volume Incremental


[psig] expelled [cm3] volume expelled
[cm3]

0 0 0

2000 7.30 7.30

4000 7.85 0.55

8000 8.12 0.27

16000 8.66 0.54

The bulk volume of the sample is 650 cm3 and its porosity is 14%. Find the
average pore compressibility.

Solution

50
We find the initial pore volume:
Vp = φVb

Vp = 0.14 × 650 = 91 cm
3

The values of the cumulative volume expelled are divided by the initial pore
volume to obtain ΔVcp/Vp. This is then plotted against the applied pressure,
and the following plot is obtained:
0.1

0.08

0.06
∆Vcp
Vp
0.04

0.02

0
0 5000 10000 15000 20000

P [psig]

As can be seen, the slope becomes constant after a certain point, and the
value of the slope after that point can be considered as the average pore
compressibility. Therefore, any two data points on the straight part of this
curve can be used to find the average pore compressibility.

Using the final two data points:


0.0952 − 0.0892
cp = = 7.5 × 10−7 psig−1
16000 − 8000

3.3 Summary

Rock compressibility is the relative change in rock volume per unit change
in pressure at a constant temperature. The change in volume is caused by
external pressure (overburden pressure) and internal pressures (fluid pressure
and matrix pressure) acting on the rock. There are three main types of rock
compressibility: matrix, bulk, and pore. Finding the pore compressibility is
important in determining the reservoir performance and in understanding
hydrocarbon production. Bulk compressibility, on the other hand, is important
for subsidence studies. Rock compressibility can be determined through a
laboratory experiment which yields data on the change in fractional volume as
the applied pressure is varied.

51
A summary of rock compressibility is presented in Table 3.1.

Table 3.1: Definition of rock compressibility and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Isothermal rock c The relative Pore/formation


compressibility change in compressibility
volume per is important in
unit change determining reservoir
in pressure performance and
at a constant in understanding
temperature. hydrocarbon
production. Bulk
compressibility
is important
for subsidence
studies which have
environmental
impacts.

End of Chapter Questions

Question 3.1

Find the pore compressibility of a rock sample with a bulk volume of 25 cm3, a
porosity of 22% and whose pore volume decreases by 0.08 cm3 when a pressure
of 750 psig is applied on it.

Question 3.2

A cylindrical rock sample has a length of 6 cm, a diameter of 4 cm, and a dry
weight of 166.4 g. When fully saturated with an oil of density 0.85 g/cm3, the
weight of this sample is 176.9 g. Assuming constant temperature, find:

52
a) The porosity of the sample

b) The pore compressibility of the sample, given that 0.2 cm3 of oil leaves
the sample when a pressure of 3200 psig is applied on it

c) The amount of oil that can be extracted from the sample when a pressure
of 2000 psig is applied on it

Question 3.3

A reservoir with an outer radius of 400 m, an inner radius of 2.5 m, and a height
of 15 m experiences a drop in pressure from 6400 psig to 5150 psig. The initial
porosity of the reservoir is 17.8%. What is the porosity of the reservoir after the
pressure drop, given that the pore compressibility of the reservoir is 8.5 ×10-5
psig-1?

Question 3.4

The porosity of a rock sample is 0.155, and its bulk volume is 85 cm3 at the ground
level at ambient conditions. The porosity of this sample drops by 20% at a depth
of 12000 ft. Find the pore compressibility of this sample assuming constant
temperature, given that the overburden pressure gradient is 0.96 psig/ft.

Question 3.5

Pressure was applied on a rock sample, and the consequent amount of fluid
expelled was recorded. The following table provides the details of the recordings:

Pressure applied [psig] Cumulative volume expelled [cm3]

0 0

3000 3.65

4000 4.15

6000 4.48

10000 5.14

14000 5.80

The dry weight of this sample is 148.3 g, and its weight when saturated with an
oil of density 0.88 g/cm3 is 163.5 g. Plot the fractional volume as a function of
applied pressure and find the average pore compressibility of this sample.

53
54
Chapter 4

Permeability
The concept of permeability is similar to that of cars on highways. Imagine that
there is a highway with only one lane and there are 300 cars that need to pass
through it; the flow of cars will be difficult as there is limited space to pass through
it. However, if the number of lanes increases to five, then the flow will become
much easier. In this case, permeability is a measure of the ease with which the
cars flow on the highway. Higher permeability means easier flow which, in this
case, will happen with the five-lane highway. Conversely, the one-lane highway
has lower permeability. Similarly, in rocks, permeability is a measure of the
ease with which the fluid flows in the porous medium. Figure 4.1 represents
a schematic showing the difference between low and high permeability rocks
in a reservoir. The permeability of a rock measured when it is 100% saturated
with a single phase (water, oil, or gas) is often called “single-phase permeability”,
“absolute permeability,” or just “permeability”. If there are two fluids flowing in a
rock, then it relates to another concept known as “relative permeability,” which
will be discussed in Chapter 9. Permeability is part of RCAL and is considered as a
flow or transport property that helps in understanding the flow in the reservoir.
The concept of permeability was first introduced by French civil engineer Henry
Darcy in 1856 when he performed an experiment on sand filtrates and analyzed
the concept of permeability. Darcy’s law for a single phase (liquid) is expressed
as:

kA
q=− dP (4.1)
µL

where q is the flow rate [m3/s], k is the permeability [m2], A is the core cross-
sectional area perpendicular to the flow [m2], L is the length of the core [m], dP is
the pressure difference across the core [Pa or N/m2], and µ is the viscosity of the
injected fluid [Pa.s or N/m2.s]

This equation is the linear form of Darcy’s law for incompressible fluid, which is
discussed further in the following sections.

55
(a) (b)

Figure 4.1: Schematic showing the cross section at the micro-scale of (a) a lower permeability
rock and (b) a higher permeability rock. The fluid flow is much easier in rock (b) compared to
rock (a).

4.1 Applications of Permeability

Permeability is a parameter that describes the flow in porous media. From Darcy’s
law, we can estimate the production flow rate from the reservoir to the surface.
This can be done by determining the permeability of the reservoir through
laboratory experiments on core samples extracted from the same reservoir, as
well as determining all the parameters associated with Darcy’s law.

4.2 Validity of Darcy’s Law for Single-Phase Permeability

Darcy’s law for single-phase flow is valid under some conditions, which include:

1. The core sample used needs to be 100% saturated with a single


phase (water, oil, or gas). If the system consists of more than one
fluid, then we need to consider relative permeability, which will be
discussed in Chapter 9.

2. The flow has to be laminar. Flow can be characterized as either


laminar or turbulent. Laminar flow is defined as the “slow,” uniform flow
while turbulent flow is defined as the “fast,” chaotic flow (Figure 4.2).
In order to determine whether the flow is laminar or turbulent, a
dimensionless number is used, known as the Reynolds number. This
number is obtained from the following equation:

ρvD
Re = (4.2)
µ

where Re is the Reynolds number [dimensionless], ρ is the density of the


fluid [kg/m3], v is the velocity of the fluid [m/s], D is the pipe diameter
[m], and µ is the viscosity of the fluid [Pa.s]. A flow with Reynolds number of
2100 or less is considered laminar.

This equation is mainly used in pipes as they have a fixed diameter;


however, for a porous medium, an average grain diameter is used. Since

56
it is difficult to compute Reynolds number in porous media, another
technique is used to determine whether the flow is laminar or turbulent.
This technique is used when measuring permeability in the laboratory
and will be discussed in the following sections. It is important to mention
that in a reservoir, the flow is generally laminar.

3. The flow has to be steady-state flow. Steady-state flow means that


whatever enters the system leaves the system, or that there is no
volumetric change over time. This is true for fluid flow in core samples if
the same amount of fluid enters and leaves the system over a certain
period. If the flow is unsteady-state, then Darcy’s law is invalid.

(a) (b)

Figure 4.2: Schematic showing the two different types of flow: (a) laminar flow (a flow that is
uniform, smooth, and occurs at low flow rates) and (b) turbulent flow (a flow that is chaotic and
occurs at high flow rates compared to laminar flow). The flow in porous media has to be laminar
in order for Darcy’s law to be valid.

4.3 Darcy’s Law Under Different Boundary Conditions

In this section, we will derive Darcy’s law for different boundary conditions by
starting from the differential form of Darcy’s law:

kA dP
q=− (4.3)
µ dx

where q is the flow rate [m3/s], k is the permeability [m2], A is the cross-sectional
area perpendicular to flow [m2], dx is the change in length [m], dP is the pressure
difference across the core [Pa], and µ is the viscosity of the injected fluid [Pa.s].

Fluids can be either compressible or incompressible. “Fluid” is a term that


refers to liquids and gases. Compressible fluids are fluids that change volume
due to a change in pressure, such as gases. Liquids, on the other hand, are
considered incompressible because their volumes change negligibly with a
change in pressure. An example of compressible fluids in real-life applications
is scuba tanks, where air is compressed inside the tank. The actual volume of
the compressed air in the tank, under atmospheric pressure, is much larger
than the actual volume of the tank itself. However, if we wish to fill the same
tank with water (which is an almost incompressible fluid), the volume of water
that fills the tank will approximately be the same inside and outside the tank.
When we consider flow in porous media, we need to consider whether the flow is
compressible or incompressible, as this will change the governing equations. In

57
addition, we need to consider whether the flow is linear in Cartesian coordinates
(most laboratory experiments) or radial coordinates (reservoir conditions).

4.3.1 Case 1: Linear solution of Darcy’s Law for incompressible fluid

Let us assume we have the system shown in Figure 4.3, and we inject it with an
incompressible fluid such as water. Based on the system, we know that the length
is L, the inlet pressure is P1, and the outlet pressure is P2. In order for the flow to
occur from the inlet to outlet, P1 needs to be greater than P2, because fluid flows
from high pressure to low pressure. This is analogous to the movement of most
of the other flows, such as the flow of electrical charges from high to low voltage,
and heat transfer from high to low temperature. Our homes also have a real-life
example of such flows. In a vacuum cleaner, for the particles to flow to the dust
collector, a pressure lower than the atmospheric pressure is required inside the
appliance. The vacuum cleaner achieves this by creating a partial vacuum inside
the machine for the particles or dust to flow towards it.

Darcy’s law is usually written with a negative sign. This is because dP = P2 – P1 and
since P2 is smaller than P1, a negative sign will make the overall equation positive.
For simplicity, we will use the following differential form of Darcy’s law in this
book:

kA dP
q= (4.4)
µ dx

where dP = P1 – P2 and P1 > P2.

Now, we can rearrange the equation to obtain:

kA
qdx = dP (4.5)
µ

Taking the integral with the boundary limits as shown in Figure 4.3:
 L  P1
kA
q dx = dP (4.6)
0 µ P2

Then, the equation becomes:

kA
q (L − 0) = (P1 − P2 ) (4.7)
µ

Finally, we divide both sides by L:

kA
q= (P1 − P2 ) (4.8)
µL

58
Equation 4.8 is the final form of Darcy’s law for an incompressible linear system.

q
A

P1 dP P2

0 dx L

Figure 4.3: Schematic showing a cylindrical core sample used for permeability measurement,
where a cross section is taken to display the boundary limits of the linear system.

4.3.2 Case 2: Radial solution of Darcy’s Law for incompressible fluid

In this case, the only difference is that the flow occurs in a radial manner as
opposed to linear. This flow is more representative of a reservoir. One of the
changes is that the area perpendicular to the flow is the circumference of the
circle (2πr) multiplied by the thickness of the reservoir (h) (in contrast to πr2 in
the linear flow for a cylindrical-shaped object). The second change is that instead
of the flow varying in the x-coordinate, it will vary in the r-coordinate and thus
instead of having dP/dx, we will have dP/dr. Figure 4.4 shows the radial system
of the reservoir.

We start with the differential form of Darcy’s law (Equation 4.4) and substitute
the area perpendicular to flow, which is the circumference of the circle, and
change the coordinates from linear to radial. The equation becomes:

2kπrh dP
q= (4.9)
µ dr

59
Then, we rearrange the equation to obtain:
dr 2πkh
q = dP (4.10)
r µ

Now, we take the integral with the boundary limits in agreement with Figure 4.4,
thus obtaining:
 re  Pe
1 2πkh
q dr = dP (4.11)
rw r µ Pwf

where re is the reservoir’s outer radius [ft], rw is the well radius [ft], Pe is the
reservoir’s outer pressure [psia], and Pwf is the wellbore flowing pressure [psia].

Then, we integrate the equation with the given boundary condition to obtain:

2πkh
q (ln re − ln rw ) = (Pe − Pwf ) (4.12)
µ
Finally, we rearrange the equation and use the natural logarithm properties to
obtain:
2πkh (Pe − Pwf )
q= (4.13)
µln rrwe
This is the final form of Darcy’s law for incompressible radial system.

Dealing with gases is different from liquids as gases are compressible fluids.
Hence, we need to account for the change in volume when deriving the equation.
We will discuss that in the later sections.

In addition, equations for Darcy’s law for different systems can also be modified
to account for the gravity term in vertical systems. Nevertheless, these equations
are rarely used in petroleum engineering, as the flow in the reservoir and in
laboratory experiments is mainly horizontal. Therefore, it is not discussed in
this book.

60
re

q(r+dr)
dr

rw
Center of
the well
q(r)

Pwf dp Pe h

Pwf dr Pe h

rw
r
r + dr
re

(a) (b)
Figure 4.4: Schematic showing (a) a radial system where the flow occurs from the outer boundary
to the wellbore region and (b) a zoomed in section of one part of the reservoir to clearly explain
the system.

Unit Systems

When using Darcy’s law, three main unit systems can be used to find permeability.
These units are explained in Table 4.1. It is important to note that since
permeability values are very small in m2, a unit of Darcy [D] or milli Darcy [mD]
is used, named in honor of Henry Darcy. Moreover, the unit bbl/d in the table
indicates barrel (bbl) per day (d), which is a standard unit to quantify volume in
the oilfield units.

Table 4.1: Summary of different units used when dealing with permeability.

Parameter SI Units Darcy Units Oilfield Units

q (flow rate) [m3/s] [cm3/s] or [cc/s] [bbl/d]

L (length) [m] [cm] [ft]

A (area) [m2] [cm2] [ft2]

h (thickness) [m] [cm] [ft]

r (radius) [m] [cm] [ft]

P (pressure) [Pa] [atm] [psia]

k (permeability) [m2] [D] [mD]

µ (viscosity) [Pa.s] [cP] [cP]

61
By rearranging Equation 4.8, the following equation is obtained, which can be
used to find permeability in SI or Darcy units:

qµL
k= (4.14)
A(P1 − P2 )

However, for oilfield units, this equation is multiplied by a conversion factor for
linear flow:

qµL
k = 887.2 (4.15)
A(P1 − P2 )

and similarly, for radial flow:

qµln rrwe qµln rrwe


k = 887.2 = 141.2 (4.16)
2πh(Pe − Pwf ) h(Pe − Pwf )

Example 4.1
In an experiment designed to measure the permeability of a core sample
using brine, the cylindrical core was saturated with brine, and then water
was injected at a rate of 0.07 cm3/s. The pressure at the inlet was measured
to be 43 psig while the outlet was open to atmospheric pressure. Knowing
that the core length is 3.5 in, the core diameter is 1.48 in, and the brine
viscosity is 1 cP, calculate the permeability of the core.

Solution
The units must always be consistent when performing these calculations.
Therefore, the flow rate must be converted from cm3/s to bbl/day so that
all the variables are in oilfield units:

1 cm3/s = 0.54 bbl/d

0.07 cm3/s = 0.0378 bbl/d

Furthermore, the length and diameter of the core must also be converted
from inch to feet. This will be done in the calculation.

Equation 4.15 can be used to find permeability (in oilfield units):

62
qµL
k = 887.2
A(P1 − P2 )
0.0378 × 1 × 3.5
k = 887.2
π( 0.74 2
12
= 19.04 mD
12 ) × (43 − 0)
Note that the answer when solving for permeability in oilfield units will
always be in mD. Also note that a conversion factor of 887.2 is used when
solving in oilfield units, but it is not used when solving for permeability in SI
or Darcy units.

Example 4.2
a) Derive the conversion of permeability from D to m2

b) Convert a permeability of 3.5 mD to m2

c) Convert a permeability of 7.3×10-12 m2 to D

Solution
a) For this conversion, the unit systems in Table 4.1 need to be analyzed.
Using Equation 4.14, we obtain:
qµL
k=
A(P1 − P2 )
We now need to reduce this equation to its base units for both the Darcy
unit system and the SI unit system. For the Darcy unit system, we obtain:

3 2
cm .cP. cm cm .cP
k= 2 = =D
s. cm .atm s.atm

For the SI unit system, we obtain:


3
m .Pa.s.m
2
m .Pa.s
k= 2 = = m2
s.m .Pa s.Pa

It is apparent that a conversion factor is required to convert permeability


from D to m2. To find this conversion factor (we will call it β), the permeability
equations obtained from the two unit systems will be equated:

63
2
D =β m
−4 m2
2
cm .cP
2
m .Pa.s β=
10
×
Pa
×
1 cP

=β 2
s.atm s.Pa
m 101325 Pa 1000 cP

2
9.869 × 10−13
cm Pa cP 10
−4

β= 2 × × β= =
m atm Pa.s 101325 × 1000

Therefore, 1 D would be equal to 9.869 × 10-13 m2.

b) Using the conversion factor β:


2
1 D =β m

3.5 mD = 3.5 × 10
−3
D = 3.5 × 10
−3
×β m
2


3.5 mD = 3.5 × 10
−3
× 9.869 × 10
−13
m
2
= 3.45 × 10−15 m2
c) Using the conversion factor β:
2
1 D =β m
D
2
1 m =
β

7.4 D
7.3 × 10
−12 7.3 × 10−12
7.3 × 10
−12
m
2
= D = −13 D =
β 9.869 × 10

4.4 Laboratory Measurements of Absolute Permeability

The permeability of core samples is measured using either liquid or gas. The
procedure and the governing equation is different for each case.

4.4.1 Liquid Permeability

Before we start measuring the permeability of a core sample, we first need to


measure its dimensions (length and diameter) as length and area are part of
Darcy’s law. Secondly, we insert the core sample in a core holder as shown in
Figure 4.5. A common procedure is to vacuum the core using a vacuum pump
prior to injecting any liquid. This is to remove any air from the system and to
ensure the flow of only one phase. Then, a liquid (in this case water) is injected
at a specific rate using a pump. Before injecting the water, we need to ensure
that a confining pressure, similar to the overburden pressure that is squeezing
the core sample from all the sides, is applied. This is done in order to guarantee
that the water is only going through the core sample and not bypassing it as that
will generate errors in the measurement. A general rule of thumb is to apply a
confining pressure (the oil pressure in Figure 4.5) that is at least 1.5 times higher
than the water injection pressure. When water is injected at a constant flow rate,

64
a wait time is required in order to achieve a steady- state flow where the inlet and
outlet pressures become constant and do not fluctuate. After the steady-state
flow is achieved, we record the flow rate, inlet pressure, and outlet pressures. We
then move on to a new flow rate and follow the same procedure. After recording
a few data points, we can plot the data in order to find the permeability of the
core sample.

If we rearrange Darcy’s law for the liquid phase (Equation 4.8), it becomes:

q k dP
= (4.17)
A µ L

This equation is in the linear form y = mx+b, where y is the dependent variable
(q/A), x is the independent variable (dP/L), m is the slope (k/µ), and b is the
y-intercept. The y-intercept in this case is 0 since the curve goes through the
origin.

If we plot this figure with our data, a figure similar to Figure 4.6 will be generated.
The slope of that figure will be equivalent to the permeability divided by the
viscosity. In order to obtain the permeability, the slope needs to be multiplied
by the viscosity. When analyzing experimental data, make sure you follow
consistent units as shown in Table 4.1. Darcy’s units are the most commonly
used units when analyzing laboratory data.

0 00
30

Oil

0 00
30
0 00
30

Water
Core

Figure 4.5: Schematic showing the experimental set-up to measure the liquid permeability of
a core sample. In this system, we inject water through the core sample and use oil to apply a
confining pressure, which should be higher than the water injection pressure.

65
q
A

k
Slope=
μ

k = (slope . μ)
(P1 - P2)
L

Figure 4.6: A plot analyzing laboratory measurements of liquid permeability. The y-axis is q/A and
the x-axis is dP/L while the slope is equivalent to k/µ.

4.4.2 Gas Permeability

Dealing with gas is different because gas is compressible as opposed to liquids.


Thus, the governing equations will be different as well to account for this change
in volume. Measuring the gas permeability has advantages over measuring liquid
permeability. The measurements for gas take less time than those for liquid,
and gas does not wet the core, which means the core can be reused for further
analyses. The disadvantage is that gas permeability requires correction as it
tends to be overestimated compared to liquid permeability. Figure 4.7 shows the
typical experimental set-up used to measure gas permeability. Since it is difficult
to control the gas flow rate, the core holder is attached to a gas cylinder, and a
metering valve is used to vary the flow rate. There is a flow meter at the end of
the core to measure the flow rate at atmospheric conditions. The simplest way
to derive the equation for the gas permeability is to take the average pressure
across the core:

P 1 + P2
P̄ = (4.18)
2
where P is the average pressure across the core [atm].

We use the average pressure because the flow rate varies across the core and
an average value would be more representative of the flow in the core. We can
use Boyle’s law as shown in Equation 2.11. We divide both volumes by time (t) in
order to convert it to flow rate. Thus, the equation becomes:

qa Pa = q̄ P̄ (4.19)

66
where qa is the atmospheric flow rate [cm3/s], Pa is the atmospheric pressure [1
atm], and q is the average flow rate across the core [cm3/s].

We compare the average flow rate across the core to the flow rate at atmospheric
conditions; since the atmospheric pressure is 1 atm, it can be eliminated from
the equation.

If we rearrange Equation 4.19, it becomes:

q a Pa
q̄ = (4.20)

Now, we substitute this equation in Equation 4.8:

kA qa Pa
q̄ = (P1 − P2 ) = (4.21)
µL P̄

Next, we can move to the other side of the equation and substitute Equation
4.18 in it to obtain:

kA (P1 + P2 )
qa = (P1 − P2 ) (4.22)
µPa L 2

Then, by rearranging the equation and factorizing it, we obtain:

kA  2 
qa = P1 − P22 (4.23)
2µPa L

For simplicity, Pa is substituted as 1 atm when using Darcy’s units and thus the
equation becomes:

kA  2 
qa = P1 − P22 (4.24)
2µL

Similar to liquid permeability, we can rearrange the equation to find the gas
permeability across the core after acquiring several data points (Figure 4.8).
When dealing with gases, it is more common to reach a higher flow rate than
liquids, because gases have lower viscosity that can result in turbulent flow,
making Darcy’s law invalid. This can be seen from the plot. A laminar flow, which
might also be referred to as “Darcy’s flow,” will follow the slope from the origin
and, once the slope deviates, enter the turbulent flow regime or “non-Darcy’s
flow”. The data that falls in the turbulent flow regime has to be omitted
from the analysis.

When comparing gas permeability to liquid permeability, gas permeability


tends to be higher than the latter. This is due to a gas slippage at the pore walls
known as the Klinkenberg effect. This gas slip makes the permeability higher

67
than what it should be; therefore, it is not representative of the actual value.
Fortunately, this can be corrected by computing the gas permeability (kg) at every
data point and then plotting it against the inverse of P, as shown in Figure 4.9.
The y-intercept of this line is the equivalent liquid permeability (kL). The x-axis
of the plot is the inverse of P, so a zero value of x represents infinite pressure.
At infinite pressure, gas can be considered to behave as a liquid. One factor
that affects this slippage is gas molecular weight. As the gas molecular weight
increases, the slippage decreases since gas becomes heavier and closer to liquid.
This effect is shown in Figure 4.10.

0 00
30

Oil

0 00 0 00
30 30

Gas
Core
Metering Valve
Flow Meter

Gas Cylinder

Figure 4.7: Schematic showing the experimental set-up for measuring the gas permeability of a
core sample.

Laminar Flow Turbulent Flow

qa
A

k
Slope=
μ

k = (slope . μ)
(P12 - P22)
2L

Figure 4.8: Analysis of laboratory measurements of gas permeability using Pa = 1 atm.

68
Slippage

kg

Absolute or Liquid Permeability (kL)

1
P

Figure 4.9: A plot showing how to correct the gas permeability to liquid permeability through
plotting the gas permeability (kg) of each data point as a function of the inverse of the average
pressure. The intercept of this line is the corrected liquid permeability (kL).

Increasing gas molecular weight

He

N2

CO2
kg

He < N2 < CO2

1
P

Figure 4.10: A plot showing the effect of gas molecular weight on gas slippage. The higher the gas
molecular weight, the lower the slippage.

69
Example 4.3
A core was mounted in a gas permeameter to measure permeability.
Determine the liquid permeability of the core using gas (in Darcy’s units).
The laboratory data are as follows:

- Diameter of the core = 3.78 cm


- Length of the core = 7.63 cm
- Gas viscosity = 0.0148 cP

q [cm3/s] P1 [psig] P2 [psig]

9.07 6.45 1.27

3.44 2.67 0.49

1.41 1.13 0.22

Solution
First, we will ensure that the units are consistent. We will use Darcy units,
so the dimensions will be in cm, the viscosity will be in cP, the flow rate will
be in cm3/s, and the pressures will be in atm. The cross-sectional area of the
core will first be calculated:


d2 3.782 2
A=π =π = 11.22 cm
4 4

Next, the pressures will be converted from psig to atm. The table will then
look like the following:

q [cm3/s] P1 [atm] P2 [atm]

9.07 1.439 1.086

3.44 1.182 1.033

1.41 1.077 1.015

Since we are dealing with gas permeability, Equation 4.24 will be rearranged
so that the permeability for each flow rate can be calculated:

2qµL
kg =
A(P12 − P22 )

70
Furthermore, the average pressure for each flow rate will also be calculated:

P 1 + P2
P̄ =
2
The permeability and average pressure for each flow rate is calculated and
recorded in the table below, along with the inverse of the average pressure:

kg [D] P [atm] 1⁄P [atm-1]

0.204 1.262 0.792

0.213 1.107 0.903

0.219 1.046 0.956

The permeability is then plotted against the inverse of the average pressure,
and the y-intercept of the line of best fit through the data point gives the
corrected liquid permeability of the core:
0.3

0.25

0.2

kg [D] 0.15
y= 0.0923x+0.1305

0.1

0.05

0
0 0.2 0.4 0.6 0.8 1
1
[atm-1]
P

From the equation of the line, the corrected liquid permeability of the core
is 0.1305 D.

4.5 Pressure Profile

By understanding rock permeability, we will be able to analyze the pressure


profile across core samples. The pressure profile differs depending on the
injected phase, and so the governing equations are different for gas and liquid.

71
4.5.1 Pressure Profile: Liquid Flow

By knowing the permeability, we can estimate the pressure at any point in the
core. This can be done by rearranging Darcy’s law for liquid (Equation 4.8) and
by considering the outlet pressure P2 as the varying parameter P(x). The resulting
equation becomes:

P (x) = P1 − x (4.25)
kA
By analyzing this equation, we can expect the pressure profile to be linear.
Furthermore, by varying the distance across the core, we can find the pressure
across the core until we reach its end, which will represent P2. Figure 4.11 shows
the concept of pressure profile when using a liquid.

P1 P(x1) P2

P1

P(x1)
P

P2

0
0 x1 L
x

Figure 4.11: Pressure profile for a core sample when using a liquid.

4.5.2 Pressure Profile: Gas Flow

Since dealing with gases is different from dealing with liquids, we have to use
Equation 4.23 for this analysis. Again, we will replace P2 by P(x) and rearrange
the equation to obtain:

2qa µ
P 2 (x) = P12 − x (4.26)
kA

72
Then, we will take the square root for both sides to obtain:

2qa µ
P (x) = P12 − x (4.27)
kA
For this equation, the pressure profile would have a parabolic shape as shown
in Figure 4.12.

The concept of pressure profile can go beyond core samples and can be extended
to analyze reservoirs. It is very important to understand the pressure profile as it
can help in understanding fluid flow in reservoirs.

P1 P(x1) P2

P1

P(x1)

P2

0
0 x1 L
x

Figure 4.12: Pressure profile for a core sample when using a gas.

73
Example 4.4
The pressure profile inside a core sample is given below, where nitrogen of
viscosity 0.0178 cP flows at a constant rate. Knowing that the permeability
of the rock is 12.4 mD, the length is 6 in, and the diameter is 1in,
what is the expected gas flow rate if the flow is in accordance with
Darcy’s law?
350

300

250

200
P [psig]
150

100

50

0
0 1 2 3 4 5 6 7
Distance from inlet [in]

Solution
It is important to ensure that all the units are consistent when solving such
problems. In this case, all the given units will be converted to Darcy units
in order to obtain the flow rate in cm3/s. This means that the length and
diameter need to be converted from inches to cm, and the pressure needs
to be converted from psig to atm.

Initially, a point x on the pressure profile will be selected, and its pressure
and distance will be recorded. At a distance of 6 in, the pressure is recorded
to be 75 psig from the pressure profile.

Now, all the units will be converted to Darcy units:

Permeability = 12.4 mD = 0.0124 D


Length = 6 in = 15.24 cm
Diameter = 1 in = 2.54 cm
Distance, x = 6 in = 15.24 cm
P(x) = 75 psig = 5.1 + 1 = 6.1 atm
P1 = 300 psig = 20.41 + 1 = 21.41 atm

Note that for the pressure to be converted to atm, 1 atm needs to be added
since we have gauge pressure (psig) and not absolute pressure (psia).

74
Now, the cross-sectional area of the sample will be calculated:
d2 2.542
A=π =π = 5.07 cm
2
4 4
Finally, Equation 4.26 can be rearranged to find the gas flow rate:
2qg µ
P (x)2 = P12 − x
kA

qg =
kA(P12 − P (x)2 )
2µx
=
0.0124 × 5.07 × (21.412 − 6.102 )
2 × 0.0178 × 15.24
= 48.8 cc/s

4.6 Flow in Layered Systems

The objective of understanding the flow in a layered system is to find the average
permeability across that system with beddings of different permeability. The
concept is similar to an electrical circuit; the average permeability will vary if the
flow is in parallel or series to the beddings in the system. In addition, it could also
vary if the flow is linear or radial. We will now examine each case individually.

4.6.1 Case 1: Linear Flow in Parallel

Let us assume a system similar to Figure 4.13 which has three beddings with
different permeabilities parallel to each other, and we try to find the average
permeability in that system. First, we can see a constant pressure difference
across the system; however, the flow rate will be different across each layer as
the flow rate is a function of the permeability of that bedding. Therefore, we can
say that the total flow rate (q) is equal to:

q = q1 + q2 + q3 (4.28)

We also know that the summation of the thickness of each layer is equal to the
total thickness of the system:

h = h1 + h2 + h3 (4.29)

The total flow rate of the system is equal to:

k̄W h (P1 − P2 )
q= (4.30)
µL

75
Then, we substitute for the flow rate from Darcy’s law in Equation 4.28:

k̄W h (P1 − P2 ) k1 W h1 (P1 − P2 ) k2 W h2 (P1 − P2 ) k3 W h3 (P1 − P2 )


q= = + +
µL µL µL µL
(4.31)

Now, we can factor the common parameters in the equation which will lead to:

k̄h = k1 h1 + k2 h2 + k3 h3 (4.32)

Finally, we can generalize the equation to obtain:


n
 k i hi
k̄ = (4.33)
i=1
h

This equation can be used to estimate the average permeability in a linear system
where the beddings are parallel to each other.

q1
P1 P2
q2
k1 h1

q
k2 h2 q
q3 h

k3 h3
W

Figure 4.13: Schematic showing a linear system with parallel beddings of different permeabilities.

4.6.2 Case 2: Linear Flow in Series

Let us assume a system shown in Figure 4.14 which has three beddings in series
with each other; each bedding has a different permeability. Here, the same flow
rate passes through all these rocks:

q = q 1 = q2 = q 3 (4.34)

However, the pressure across each rock is different, as shown in Figure 4.14 and
explained in Figure 4.15. It can be said that:

P1 − P2 = ∆P1 + ∆P2 + ∆P3 (4.35)


76
Furthermore, the total length is composed of the length of each rock:

L = L1 + L2 + L3 (4.36)

We know that the flow rate across the entire system is equal to:

k̄W h (P1 − P2 )
q= (4.37)
µL

Now, we will substitute Darcy’s law in Equation 4.35 by making ∆P of each


respective equation as the subject to obtain:

qµL qµL1 qµL2 qµL3


P1 − P2 = = + + (4.38)
k̄W h k1 W h k2 W h k3 W h
We can now remove the common parameters and the equation becomes:

L L1 L2 L3
= + + (4.39)
k̄ k1 k2 k3

We can also represent this equation in the following general form:

L
k̄ = n Li
(4.40)
i=1 ki

This equation can be used to estimate the average permeability in a linear


system, where the beddings are in series with each other.

P1 P2

k1 k2 k3
q q

P1 P2 P3
h
L1 L2 L3

Figure 4.14: Schematic showing a linear system with beddings of different permeabilities
in series.

77
P1 P2
k1 > k2 > k3

k1 k2 k3

P1

k
P

P2

0
0 L
x

Figure 4.15: Schematic showing the pressure profile across a composite system with beddings in
series with each other.

Example 4.5
A rock consists of three layers of different permeabilities. The geometry
of the rock is given in the figure below, along with the permeability and
dimensions of each layer. Find the average permeability of the rock.

1.7 mD 0.5 in
q 32.5 mD
1 in
5.3 mD

3 in 5 in

Solution
The three layers given in this question can be numbered as follows: top-
left is layer 1, bottom-left is layer 2, and right is layer 3. To find the average
permeability of the entire rock, the combined permeability of layers 1 and 2
can first be found so that the combined section can be treated as one layer.

78
This combined layer along with layer 3 can be used to find the overall
permeability.

Since the flow is horizontal, layers 1 and 2 are parallel to each other. The
combined parallel permeability of layers 1 and 2 can be calculated using
Equation 4.33:
n
 k i hi
k̄ =
i=1
h


k 1 h1 + k 2 h 2 1.7 × 0.5 + 5.3 × 0.5
k1,2 = = = 3.5 mD
h 0.5 + 0.5
Layers 1 and 2 combined are in series with layer 3. Therefore, the average
permeability can be calculated using Equation 4.40:
L
k̄ = n
Li
i=1 ki

k̄ =
L1,2 + L3
L1,2 L3
= 3
3+5
5 =
+ 32.5
7.91 mD
k1,2 + k3 3.5

4.6.3 Case 3: Radial Flow in Parallel

The flow in parallel systems in radial orientation (Figure 4.16) is similar to that in
linear orientation, as will be proven below.

First, the overall flow rate is the summation of all the flow rates across the layers
(Equation 4.28), and the total thickness is the summation of all the thicknesses
across all the layers (Equation 4.29).

The total flow rate for this system is equivalent to:

2π k̄h (Pe − Pwf )


q=   (4.41)
µln rrwe

Now, if we substitute each flow rate in each layer in Equation 4.28, we will have:

2π k̄h (Pe − Pwf ) 2πk1 h1 (Pe − Pwf ) 2πk2 h2 (Pe − Pwf ) 2πk3 h3 (Pe − Pwf )
  =   +   +  
µln rrwe µln rrwe µln rrwe µln rrwe
(4.42)

We can now factor the common parameters to obtain:

k̄h = k1 h1 + k2 h2 + k3 h3 (4.43)

79
The form is indeed the same as the linear system and can also be represented in
the same general form:
n
 k i hi
k̄ = (4.44)
i=1
h

This equation can be used to estimate average permeability in a radial system


where the beddings are parallel to each other.

re
rW

k1 h1 q1

k2 h2 q2

q3
k3 h3

Figure 4.16: Schematic showing a radial system with beddings of different permeabilities in
parallel.

4.6.4 Case 4: Radial Flow in Series

The flow in series for a radial system is shown in Figure 4.17. We know that the
flow rate across the layers is the same, as shown in Equation 4.34; however, the
pressure difference in each layer is different as shown in Equation 4.35. The flow
rate in a radial system is expressed in Equation 4.41.

By substituting Darcy’s law in Equation 4.41 and making ∆P of each respective


equation as the subject, we obtain:
     
re r1 re
qµ ln rw qµ ln rw qµ ln r1
Pe − Pwf = = + (4.45)
2π k̄h 2πk1 h 2πk2 h

80
Then, we eliminate the common parameters and rearrange the equation to
obtain this generic form:
 
re
ln rw
k̄ = r  (4.46)
n ln
(i+1)
ri
i=1 ki

where r(i+1) represents the outer radius of layer i and ri represents the inner layer.

This equation can be used to estimate the average permeability in a radial


system where the beddings are in series with each other.

re

r1

rW

h k2 k1

Figure 4.17: Schematic showing a radial system with beddings of different permeabilities in
series.

Example 4.6
A radial system consists of three layers with the following properties:

- Layer 1 has an inner radius of 0.25 ft, an outer radius of 5 ft,


and a permeability of 10 mD
- Layer 2 has an inner radius of 5 ft, an outer radius of 150 ft,
and a permeability of 80 mD
- Layer 3 has an inner radius of 150 ft, an outer radius of 750 ft,
and a permeability of 150 mD

81
Calculate the average permeability of this system if the flow is in series.

Solution
In this radial system, the flow is in series as it moves from one layer to the
other from the inside out, as shown in Figure 4.17. As such, the equation for
calculating average permeability for radial flow in series will be used.

Using Equation 4.46:


 
re  750 
ln
k̄ =
n ln
rw
r
(i+1)
 = 5
ln( 0.25 )
ln 0.25
ln( 150
5 )
ln( 750
150 )
= 22.7 mD

ri
10 + 80 + 150
i=1 ki

4.7 Flow in Channels and Fractures

The concept of permeability can go beyond core samples and beddings, and can
be useful in estimating permeability in channels and fractures. Channels and
fractures can be superficially induced in the reservoir in order to increase the
permeability. However, some reservoirs can be naturally fractured.

4.7.1 Flow in Channels

Channels are created in reservoirs by injecting acids to dissolve the rock in


order to increase the permeability in the near-wellbore region. Increasing the
permeability will increase the flow rate of the hydrocarbons in the well, and will
hence increase the productivity. Channels can have different shapes; however,
the simplest one is the capillary tube shape (Figure 4.18), which also represents
the wormhole effect caused by acidizing.

The flow in a capillary tube is described by Poiseuille in the following equation:

πr4
q= (P1 − P2 ) (4.47)
8µL

We know that the area of the channel is equal to πr2; thus, we can write:

Ar2
q= (P1 − P2 ) (4.48)
8µL

82
Now, if we compare this equation with Darcy’s linear flow of liquids, we obtain:

kA Ar2
(P1 − P2 ) = (P1 − P2 ) (4.49)
µL 8µL

We can now eliminate all the common parameters to obtain:

r2
k= (4.50)
8
This represents a measure of permeability in channels. However, you need
to bear in mind that the channel’s radius would have a unit other than Darcy
(usually inches), and therefore unit conversion is required. If the porous medium
has n number of identical channels, then Equation 4.50 can be modified to:

r2
k=n (4.51)
8

P2

P1
l
ne
an
Ch

Matrix L

q
Figure 4.18: Capillary tube-shaped channels.

4.7.2 Flow in Fractures

Fractures can be either natural or induced, and the simplest model assumes
a slab of constant thickness as shown in Figure 4.19. The flow in this slab is
described by Buckingham in the following equation:

Ah2
q= (P1 − P2 ) (4.52)
12µL

Now, if we compare this equation with Darcy’s linear flow of liquids, we obtain:

kA Ah2
(P1 − P2 ) = q = (P1 − P2 ) (4.53)
µL 12µL

83
We can now eliminate all the common parameters to obtain:

h2
k= (4.54)
12
This should represent the permeability of fractures with a simplified shape.
Nevertheless, we still need to be careful when dealing with units.

W
L
Figure 4.19: Schematic showing a porous medium with a fracture.

Example 4.7
a) Find the fracture permeability (in D) of a cubic rock with a fracture of
width 0.5 inch passing completely through the rock.

b) Given that the rock has sides of 1.5 ft, a matrix permeability of 2.5 mD
and the flow is normal to the direction of the fracture, find the average
permeability of the rock (in mD).

Solution
a) Since the fracture permeability formula is in terms of SI units, the units
given in the question have to be converted to SI. Since 1 inch = 0.0254 m,
the fracture height, h, is:

h = 0.5 in = 0.5 × 0.0254 = 0.0127 m

Using the fracture permeability formula (Equation 4.54):

84
h2 0.01272
k= = = 1.344 × 10
−5 2
m
12 12
Using the conversion of Darcy to m2 derived in Example 4.2, the fracture
permeability will be 1.36 × 107 D.

b) The fracture that passes through the length of the rock essentially divides
it into three sections. There is a matrix section above and below the fracture
with permeabilities of 2.5 mD, and then there is the fracture section itself
with a permeability of 1.36 × 107 D or 1.36 × 1010 mD. The exact position of
the fracture on the cube is not given, so it can be assumed that the fracture
is in the exact middle of the cube. The exact position of the fracture is
irrelevant, since the matrix will occupy the same area and hence the average
permeability would be the same regardless of where the fracture actually
lies. Since we assume that the fracture is in the middle, and the fracture has
a height of 0.5 in or 0.5/12 = 0.04 ft, then the length of each of the matrix
sections on either side of the fracture becomes:

1.5 − 0.04
L= = 0.73 ft
2
Since the flow is normal to the fracture, this becomes a case of linear
flow with three sections in series to one another. Therefore, the average
permeability of the rock can be found using Equation 4.40:


L
k̄ = n Li
i=1 ki

k̄ =
Lm + Lf + Lm
Lm Lf Lm
=
0.73 + 0.04 + 0.73
0.73 0.04
+ 1.36×10 0.73 = 2.57 mD
+ + 2.5 10 + 2.5
km kf km

As can be observed from this question, the average permeability of the rock
increases due to the fracture, since the fracture is basically empty space
through which a fluid can pass.

85
4.7.3 Average Permeability with Channels and Fractures

In the case where there are channels and/or fractures in a rock and the average
permeability needs to be found, area ratios can be used, as shown in the equation
below:
n
 k i Ai
k̄ = (4.55)
i=1
A

where Ai is the cross-sectional area of the channel, fracture, or surface [ft2] and A
is the total area [ft2].

As evident from this equation, the very high permeability of the channel or
fracture will cause the average permeability of the rock to increase significantly
even though they occupy a small area on the rock.

Example 4.8
A cube-shaped rock with sides of length 2.5 ft has a matrix permeability of
5 mD and a 0.3 inch diameter cylindrical channel that traverses the rock.

a) Calculate the permeability of the channel (in D).

b) Calculate the average permeability of the rock (in mD) when flow
is linear, and in the direction of the channels.

Solution
a) Since the channel permeability formula is in terms of SI units, the units
given in the question have to be converted to SI. Since 1 in = 0.0254 m, the
channel radius, r, is:

r = 0.15 in = 0.15 × 0.0254 = 0.00381 m

Using the channel permeability formula (Equation 4.50):

r2 0.003812
k= = = 1.815 × 10
−6 2
m
8 8
Using the conversion of Darcy to m2 derived in Example 4.2, the channel
permeability is 1.84 x 106 D.

86
b) To find the average permeability, Equation 4.55 can be used. However,
first, the area of the matrix and the channel needs to be calculated.

The area of the channel is:


d2 0.32 −4

2
Ac = π =π = 4.909 × 10 ft
4 4

Note that the diameter is converted from inches to feet in the equation.

Now, the area of the matrix is:

Am = At − Ac = 2.52 − 4.909 × 10−4 = 6.2495 ft


2

Note that to find the surface area of the matrix, the surface area of the
channel has to be subtracted from the surface area of the cube.

Now, Equation 4.55 can be used to find the average permeability:


n
 k i Ai
k̄ =
A
i=1

which is equal to:

5 × 6.2495 + 1.84 × 109 × 4.909 × 10−4


= 1.45 × 105 mD
k m Am + k c Ac
=
Am + Ac 6.2495 + 4.909 × 10−4

Note that the channel permeability is converted from D to mD to be used


in this equation.

4.8 Summary

Permeability is part of RCAL and is a measure of the ability of a porous rock


to allow fluids to pass through it, and it can therefore help determine the rate
at which oil and gas will flow from the reservoir to the surface. Darcy’s law of
single-phase permeability is the equation used to determine permeability when
only one phase (water, oil, or gas) is present, and it is only valid when the core
sample is 100% saturated with the single phase and the flow is laminar and
steady-state. Darcy’s law for incompressible fluids has a solution for both linear
and radial systems. The three main unit systems used in Darcy’s equation are
SI units, Darcy’s units, and Oilfield units, where the unit for permeability is m2,
D (Darcy), and mD (milli Darcy), respectively. It is imperative that the units are
consistent when solving for permeability. The permeability of core samples
can be measured experimentally using either liquid or gas. However, when

87
measuring permeability using gas, a correction is required due to gas slippage
at the pore walls known as the Klinkenberg effect, which makes the permeability
higher than it should be. The pressure profile across a core sample when using
liquid is linear, whereas the pressure profile when using gas is parabolic. Average
permeability can be found for a system with layers of different dimensions and
permeabilities in series or parallel with one another depending on the direction
of fluid flow, where the orientation of the layers can be either linear or radial.
Permeability can also be calculated for channels and fractures using different
equations.

A summary of permeability is presented in Table 4.2.

Table 4.2: Definition of permeability and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Permeability k A measure of the Permeability is


ease with which important in
fluid flows through determining the
a porous rock. rate at which oil
and gas will flow
from the reservoir.

End of Chapter Questions

Question 4.1

A core of length 5 cm and diameter 3 cm is saturated with brine of viscosity 1


cP. Water is injected into this core at a rate of 0.13 cm3/s. The gauge pressure at
the inlet was measured to be 180 kPa while the outlet was open to atmospheric
pressure. Calculate the permeability of this core.

a) In m2

b) In mD

88
Question 4.2

A steady-state liquid permeability test on a core sample from a reservoir of


interest was conducted. The laboratory measured data are given below:

- Plug diameter: 2.52 cm

- Plug length: 2 cm

- Test oil viscosity: 1.82 cP

- Outlet pressure: 1 atm

- Inlet pressure: 2 atm

- Flow rate: 0.275 cm3/s

Calculate the permeability of the core.

Question 4.3

Measure the brine absolute permeability for a core sample (L= 76.38 mm, D=
37.85 mm). The rock is cylindrical, and the brine viscosity is 0.001048 Pa.s. The
rest of the required data are given in the table below.

q [m3/s] P1 [psig] P2 [psig]

9.97 × 10-9
5.7 0.8

3.53 × 10-8
16 1.1

8.33 × 10-8 35.5 1.1

89
Question 4.4

A core was mounted in a gas permeameter to measure permeability, and the


following laboratory data are obtained from the experiment:

- Diameter of the core = 2.59 cm

- Length of the core = 6.03 cm

- Gas viscosity = 0.0148 cP

q [cm3/s] P1 [kPa] P2 [kPa]

13.09 65.4 13.5

7.91 44.6 11.2

4.12 25.3 6.4

3.23 19.4 4.1

Determine the gas permeability and equivalent liquid permeability of the core
using the above data (in Darcy units).

Question 4.5

A rock sample has a length of 20.32 cm and a diameter of 5 cm. The pressure
profile inside the core sample is given below, where nitrogen of viscosity 0.0178
cP is flowing at a constant rate of 55 cm3/s. What is the permeability of the sample
if the flow is in accordance with Darcy’s law?

450
400
350
300
250
P [psig]
200
150
100
50
0
0 1 2 3 4 5 6 7 8 9
Distance from inlet [in]

90
Question 4.6

Nitrogen of 0.0178 cP viscosity is injected at an inlet pressure of 120 psig with


the outlet pressure set at 72 psig. The rock has a 1 in diameter, 4.5 in length, and
a permeability of 2.3 mD. Provide an accurate plot of the pressure profile inside
the rock.

Question 4.7

Rock samples 1, 2, 3 and 4 have similar cylindrical diameters of 3.5 cm. The
pressure profile inside the core samples is given below, where water of viscosity
1 cP is flowing at a constant rate of 1.2 cm3/s. Find the permeabilities of all the
four rock samples.

120
Rock 1

100 Rock 2

Rock 3
80
Rock 4

P [psia] 60

40

20

0
0 5 10 15 20 25 30

Distance from inlet [cm]

Question 4.8

Consider a layered reservoir consisting of three alternating layers 1 m thick and


1 m long, with k1= 1000 mD, k2 = 100 mD, and k3= 10 mD.

a) What is the average permeability of this reservoir if the fluid is flowing


parallel to the layering (kh)?

b) What is the average permeability of this reservoir if the fluid is flowing


perpendicular to the layering (kv)?

c) What is the kv/kh?

91
Question 4.9

A rock consists of six layers of different permeabilities. The geometry of the rock
is given in the figure below. The permeability and dimensions of each layer are
given in the table below.

a) Find the average permeability of this rock for horizontal flow (kh).

b) Find the average permeability of this rock for vertical flow (kv).

c) What is the kv/kh?

qv

qh 3
2
4

5 6

Layer Permeability [mD] Thickness [in] Length [in] Width [in]

1 50 3 30 10

2 60 3.5 9 10

3 30 1.7 21 10

4 40 1.8 21 10

5 20 2.5 18 10

6 10 2.5 12 10

92
Question 4.10

A radial system consists of four layers with the following properties:

Layer Inner Radius [m] Outer Radius [m] Permeability [D]

1 0.6 4.5 0.35

2 4.5 26.0 0.21

3 26.0 95.5 0.49

4 95.5 164.5 0.24

Calculate the average permeability of this system if the flow is in series.

Question 4.11

Consider the radial flow problem shown in top view (right), where the shaded
area represents a damaged zone due to the drilling operations. Assume there
are three horizontal layers of equal heights (thickness), and the permeability of
the undamaged formation is 45 mD in layer 1, 25 mD in layer 2, and 30 mD in
layer 3. Further assume that the radius of damage is 2.2 ft and the permeability
of the damaged zone is 1.5 mD in all the three layers. The radius of the wellbore
is 0.6 ft, and the radius of the reservoir is 450 ft.

a) What is the average permeability in each horizontal layer?

b) What is the average permeability of the three layers?

rd
rw

re

(Top View)

93
Question 4.12

A cube of reservoir rock with sides of 2.5 ft has a single vertical fracture of width
0.35 that passes completely through the rock. The matrix permeability is 1 mD.

a) What is the permeability of the fracture (in D)?

b) What is the average permeability of the cube if the flow is in the


direction parallel to the fracture (in mD)?

c) What is the average permeability of the cube if the flow is in the


direction normal to the fracture (in mD)?

Question 4.13

A cube-shaped rock with sides of length 1.2 m has a matrix permeability of 0.8
mD and four 0.65 cm diameter cylindrical channels that traverse the rock as
shown below. Calculate the average permeability of the rock (in mD) when flow
is linear and in the direction of the channels.

Question 4.14

The rock sample shown below has a channel that partially traverses the length
of the sample. The rock sample had a permeability of 8 mD before the channel
was drilled into it.

3 in

Channel diameter
1 in
1 mm

6 in

a) Calculate the permeability of the channel (in D).

b) Find the average permeability of the rock sample if the flow is linear
and in the direction of the channel (in mD).

94
95
96
Chapter 5

Fluid
Saturation
The concept of fluid saturation can be explained through a glass filled with
different fluids (Figure 5.1). Figure 5.1a shows a glass filled entirely with water,
meaning that the entire pore volume (glass volume) is occupied by water and
thus the water saturation is 100%. Fluid saturation is the volume of a particular
fluid in a rock sample divided by the pore volume. In Figure 5.1b, an identical
glass is occupied by both water and oil. Here, the water saturation cannot be
100% as the oil is sharing some space with the water. Finally, Figure 5.1c has
water, oil, and gas in the same glass, and thus the saturation of each fluid will be
less than 100%.

20% Air
50%
Oil
30% Oil
100%
Water

50% 50%
Water Water

(a) (b) (c)


Figure 5.1: Schematic showing identical glasses filled with different fluids: (a) the glass is
filled with 100% water, representing Sw = 1, (b) the glass is filled with 50% water and 50% oil,
representing Sw = 0.5 and So = 0.5, and (c) the glass is filled with 50% water, 30% oil, and 20% air
(gas), representing Sw = 0.5, So = 0.3, and Sg = 0.2.

Similarly, rocks are filled with one or more fluids. Fluid saturation helps us
quantify the amount of hydrocarbons or water in the rock. We can classify the
saturation into three categories: water, oil, or gas. Water saturation, Sw, is the
volume of water in a rock divided by the pore volume:

Vw
Sw = (5.1)
Vp

97
where Sw is the water saturation [dimensionless], Vw is the volume of water in the
pore spaces [cm3], and Vp is the pore volume [cm3].

Similarly, oil saturation, So, is the oil volume divided by the pore volume:

Vo
So = (5.2)
Vp

where So is the oil saturation [dimensionless], Vo is the volume of oil in the pore
spaces [cm3], and Vp is the pore volume [cm3].

Finally, gas saturation, Sg, is the gas volume in a rock divided by the pore volume:

Vg
Sg = (5.3)
Vp

where Sg is the gas saturation [dimensionless], Vg is the volume of gas in the pore
spaces [cm3], and Vp is the pore volume [cm3].

The summation of saturations of all the fluids in a reservoir has to be 1 as the


pores have to be filled with at least one fluid. If a reservoir contains water, oil,
and gas, then the equation becomes:

Vw + V o + Vg
= Sw + So + Sg = 1 (5.4)
Vp

However, if a reservoir only contains oil and water, then Equation 5.4 will reduce
to Sw + So = 1. Although Equation 5.4 is very simple, it is helpful in finding an
unknown fluid saturation mathematically.

Similar to porosity, fluid saturation is either presented as a fraction or a


percentage; however, bear in mind that it should always be used as a fraction in
calculations.

Figure 5.2 shows an example of a microscopic rock slice illustrating water, oil,
and gas saturations in the pore spaces. Similar to porosity, fluid saturation is
important to estimate the amount of hydrocarbons in a reservoir. However, it is
important to distinguish between porosity and fluid saturation. Porosity tells us
the maximum storage capacity of a medium, while fluid saturation depicts the
exact amount of fluid occupying the pore spaces of the same medium.

98
0.5 mm 0.5 mm 0.5 mm

Rock Water Rock Water Oil Rock Water Oil Gas


(a) (b) (c)

Figure 5.2: Schematic showing a cross section of a rock at the microscopic scale: (a) all the pore
spaces are filled with water (Sw = 1), (b) the pore spaces are filled with water and oil (Sw + So = 1),
and (c) the pore spaces are filled with water, oil, and gas (Sw + So + Sg = 1).

Example 5.1
A core sample with a pore volume of 15 cm3 contains water, oil, and gas. The
water volume within the sample is 6.3 cm3, while the oil volume is 5.4 cm3.
What is the gas saturation?

Solution
The water saturation can be found using Equation 5.1:

Vw 6.3
Sw = = = 0.42
Vp 15

The oil saturation can be found using Equation 5.2:

Vo 5.4
So = = = 0.36
Vp 15

The gas saturation can then be found using Equation 5.4:

S +S +S =1
w o g
S
g = 1 − Sw − So
Sg = 1 − 0.42 − 0.36 = 0.22

5.1 Measuring Fluid Saturation

Fluid saturation measurements can be classified into two types: direct and
indirect. Direct measurements include conventional core analysis techniques
such as extraction methods (retort distillation and Dean-Stark method), while
indirect measurements include electrical properties and capillary pressure,
99
which will be discussed in Chapters 6 and 8, respectively. In this chapter, we will
focus on the direct methods, which are part of RCAL.

5.1.1 Extraction Method: Retort Distillation

For the retort distillation method (Figure 5.3), a core sample is placed in a
chamber and heated to around 1100 °F (≈593 °C). This is to evaporate all the
fluids in the system (oil and water). The vapors will rise and reach a condensing
tube where cold water is being circulated. The vaporized liquids will condense
back to liquid form and will be collected in the graduated cylinder after passing
through the condensing tube. Once we have the volumes of oil and water
from this method and by knowing the pore volume of the core sample, we can
calculate the water and oil saturations using Equations 5.1 and 5.2, respectively.
The advantages of the retort distillation method are that it can directly measure
oil and water saturations, and is a relatively fast method (usually takes less than
one hour). The main disadvantage of this method is that subjecting the core to
very high temperatures can damage it, thereby preventing the core from being
used for additional experimentation and analysis.

Cooling Water In

Core Sample

Condenser

Heating Element
1000 - 1100 ºF Graduated
Cylinder
Cooling Water
Out
Figure 5.3: Schematic showing the experimental set-up for the retort distillation extraction
method.

5.1.2 Extraction Method: Dean-Stark

This method is also known as Soxhlet extraction or solvent extraction.


Chemists Dean and Stark first designed this experimental set-up in 1920. For
this experiment, a core sample is placed at the top of a solvent flask, as shown
in Figure 5.4. The solvents used are usually toluene (hydrocarbon solvent) or
a mixture of toluene and methanol. Methanol can be used in the presence of
salty water. The solvent is heated to around 230 °F (110 °C, the boiling point of
toluene), so that the water present in both the core and the solvent evaporates
when the temperature in the system exceeds the boiling point of these fluids.
The toluene vapor will strip the oil from the core and travel upward as toluene is
miscible with oil. Once the vapor goes upward, it will reach the condensing tube
with circulating cooling water. Both fluids (water and solvent) will drop down
in the graduated cylinder. Since water is denser than the solvent, it will settle
at the bottom of the graduated tube, while the condensed solvent (being less
dense) will accumulate on top of the water until it drops down to the solvent
100
flask. This method can measure the volume of water directly, and the water
saturation can be calculated using Equation 5.1 if the pore volume of the core
sample is known. Oil volume needs further calculations using material balance,
which will be discussed in the following section. We cannot measure the oil
volume/saturation from the Dean-Stark method directly, as the solvent mixes
with oil to form a new fluid that has different fluid properties than the original oil.
Therefore, collecting both fluids will be impractical. However, we can calculate the
oil saturation mathematically since the summation of all saturations is equal to 1
and compare the value with the one obtained from the material balance method.
The advantage of the Dean-Stark method is that it does not damage the core,
and the core sample can be used for future analysis. The disadvantages are that
the experiment is time-consuming as it usually takes about 48 hours, and that
time required for the experiment is also a function of the permeability; the lower
the permeability, the more time is required. Furthermore, in this experiment, we
can only measure saturation of one fluid directly, which is water, unlike the retort
distillation method which measures both water and oil saturations.

To drain

Condenser

Cooling Water entering

Graduate tube

Core Sample

Mesh

Solvent

Electric Heater

Figure 5.4: Schematic showing the experimental set-up for the Dean-Stark extraction method.

Material Balance

As mentioned earlier, the Dean-Stark method can only measure the water
volume. In order to find the oil volume and saturation, we need to use
material/mass balance. The concept is the same as discussed in Chapter 2.

101
However, now we need to assume that there are two fluids in the system
(assuming no gas occupying the pores). It is important to note that we need to
record the weight of the core sample prior to the extraction process. First, we
know that the weight of the saturated core (assuming it contains oil and water)
prior to extraction is equal to:
Ws = Wd + Ww + Wo (5.5)
where Ws is the saturated weight [g], Wd is the dry weight of the sample that can
be obtained after the extraction process [g], Ww is the weight of the water in the
core sample [g], and Wo is the weight of the oil in the core sample [g].

We can say that the weights of water and oil in the core sample are equivalent to
their densities multiplied by their volumes, as shown below:

Ws = Wd + ρw Vw + ρo Vo (5.6)

where ρw is the density of water [g/cm3], Vw is the volume of water in the core
[cm3], ρo is the density of oil [g/cm3], and Vo is the volume of oil in the core [cm3].

We know that:

Vw + Vo + Vg = Vp (5.7)

where Vg is the volume of gas in the core [cm3], and Vp is the pore volume of the
core [cm3].

Since we only have oil and water, Equation 5.7 can be reduced to:

Vw + Vo = Vp (5.8)

For simplicity, let us assume:

Vo = x (5.9)

Then, we can rewrite Equation 5.8 to obtain:

Vw = Vp − x (5.10)

Then, we can substitute Equations 5.9 and 5.10 in Equation 5.6 to obtain:

Ws = Wd + ρw (Vp − x) + ρo x (5.11)

After that, we can rearrange the equation to obtain:

Ws − Wd − ρw Vp = x(ρo − ρw ) (5.12)

102
Now, we can solve for x and replace it with Vo:

Ws − Wd − ρw Vp
Vo = (5.13)
ρo − ρw

Finally, we can divide both sides by the pore volume to find the oil saturation:

Ws − Wd − ρw Vp
So = (5.14)
Vp (ρo − ρw )

We can cross-check the water saturation value obtained from the Dean-Stark
method using the following simple term:

Sw = 1 − So (5.15)

where Sw is the water saturation [dimensionless] and So is the oil saturation


[dimensionless].

It is important to mention that the use of material balance to find the fluid
saturations can go beyond the Dean-Stark method to measure fluid saturations
obtained from different experiments, as we will explore in the following chapters.

Example 5.2
A core sample containing only water (ρw = 1 g/cm3) and oil (ρo = 0.87 g/cm3)
has a 13.6% porosity, 3 inch length, and 1.5 inch diameter. Its saturated
weight was measured to be 144.3 g, and its dry weight was measured to be
133.2 g. Calculate the water and oil saturations.

Solution
First, the dimensions of the sample need to be converted to cm so that the
units are consistent. Since 1 inch = 2.54 cm:

D = 1.5 in = 1.5 × 2.54 = 3.81 cm



L = 3 in = 3 × 2.54 = 7.62 cm
Then, the bulk volume of this core needs to be calculated. Since this is a
cylinder:

Vb = πr2 L
 2
3.81
Vb = π × × 7.62 = 86.87 cm3
2

103
We now find the pore volume:

Vp = φVb

Vp = 0.136 × 86.87 = 11.81 cm3
Since this rock contains only oil and water, the oil saturation can now be
found using Equation 5.14:

Ws − W d − ρ w V p 144.3 − 133.2 − 1 × 11.81


So = = = 0.462
V (ρ
p o − ρ w ) 11.81(0.87 − 1)

The water saturation can then be found using Equation 5.15:


Sw = 1 − So

Sw = 1 − 0.462 = 0.538

Example 5.3
A Dean-Stark apparatus was used to determine fluid saturations of a
sandstone cylindrical core plug, which measures 2 cm in diameter and 4.5
cm in length. The initial weight of the core plug prior to extraction was 76.63
g. The volume of water recovered from the core plug was 1.42 cm3. The
weight of the core plug after it was dried was 73.94 g. The porosity of the
core plug was determined to be 23.1%, and the densities of reservoir water
and oil were 1.04 g/cm3 and 0.82 g/cm3, respectively. Determine the initial
fluid saturations in the core plug.

Solution
First, the bulk volume of this core needs to be calculated. Since this is a
cylinder:
Vb = πr2 L
 2
2
Vb = π × × 4.5 = 14.14 cm3
2

Now, we find the pore volume:

Vp = φVb

Vp = 0.231 × 14.14 = 3.27 cm3

104
Since the volume of water is given, the water saturation can be found using
Equation 5.1:

Vw 1.42
Sw = = = 0.43
Vp 3.27

To find the oil saturation, we need to calculate the volume of oil.

First, we calculate the weight of water:

Ww = Vw ρw = 1.42 × 1.04 = 1.48 g

We can use this to find the weight of oil using Equation 5.5:
Ws = Wd + Ww + Wo

W o = Ws − W d − W w

Wo = 76.63 − 73.94 − 1.48 = 1.21 g
Note that the assumption made here is that the weight of gas is insignificant.

The volume of oil can then be found:


Wo 1.21
Vo = = = 1.48 cm3
ρo 0.82
The oil saturation can be found using Equation 5.2:

Vo 1.48
So = = = 0.45
Vp 3.27

Thereafter, the gas saturation can simply be found using Equation 5.4:
S w + So + S g = 1

Sg = 1 − Sw − So

Sg = 1 − 0.43 − 0.45 = 0.12

5.2 Limitations of Using Extraction Methods to Evaluate Reservoir’s


Saturation

It is difficult to evaluate the reservoir’s saturation using the conventional core


analysis (extraction methods) because of two main reasons: drilling muds and
fluid properties.

105
5.2.1 Drilling Muds

When drilling a reservoir, two types of muds are usually used: water-based mud
(WBM) and oil-based mud (OBM). When either mud is used, part of the mud leaks
into the reservoir, as the reservoir is both porous and permeable. For instance,
if we use a WBM, the water will enter the permeable reservoir and push the
fluids inside the reservoir. This means that the water saturation in the reservoir
will increase in the region near the well where the mud has entered. In drilling
terminology, this region is called the near-wellbore region. The extraction of core
samples using coring tools usually occurs in the near-wellbore region where
the mud has flushed the original reservoir fluids in that region. This makes the
samples less representative of the reservoir’s initial condition. The same thing
can be said when OBM is used, as the oil saturation will be greater than the
actual oil saturation.

5.2.2 Fluid Properties

The physical state of fluids is a function of temperature and pressure. Liquids


become gases at high temperatures, and gases become liquids at low
temperatures. Similarly, liquids become gases at low pressures and gases
become liquids at high pressures. When the core is extracted from the reservoir,
it goes through a journey from the reservoir at elevated pressures (thousands
of psia) and temperatures (around 177 °F) to the surface at ambient conditions
(atmospheric pressure and temperatures of around 77 °F, depending on the
location). The decrease in temperature is usually too insignificant to affect the
state of the fluid, as the temperature difference between the reservoir and the
surface is not substantial enough to cause a significant phase change. However,
the pressure drops significantly and reaches a threshold point where the gas in
the oil escapes. This threshold pressure is known as the bubble point pressure,
and occurs when the first bubble of gas leaves the oil. At pressures lower than
the bubble point pressure, more gas will evaporate and leave the oil, leading to
a higher gas saturation. Consequently, performing the extraction methods on
these core samples will lead to significant errors in estimating the reservoir’s
original saturations.

What we deal with in a reservoir is a combined effect of drilling mud and the fluid
properties. Figure 5.5 explains the effect of drilling mud and fluid properties
on fluid saturation from the reservoir to the surface. Initially, we have high oil
saturation in the reservoir. Then, after drilling with a WBM, the water will invade
the pore spaces to reduce the oil saturation. Then, when we extract the core,
the core undergoes pressure and temperature changes that will change its
fluid properties. Reducing the pressure from thousands of psia to atmospheric
pressure, when the core reaches the surface, will release plenty of gas dissolved
in oil, which will reduce the oil saturation even further. The purpose of this
section is to understand why using the extraction method directly on reservoir
rocks might not produce the most accurate results. In the following chapters, we
will discuss alternative methods to estimate fluid saturations in the reservoir.

106
Despite the associated errors, these extraction methods are still in use, as they
provide a good, fast, and cheap estimate of fluid saturations within a reservoir.
This can be useful in making decisions related to the amount of hydrocarbons
present in the reservoir.

Fluid Contents
At Surface
outside of
arrow
4 e
Oil Gas Water

c
15 40 45

rfa
At Su
Gas
Expansion
Journey to
Surface 3

Expansion
Shrinkage

Expulsion
2 In reservoir
After Flushing
with water-based mud
20 ― 80

1 In reservoir Original Fluids 70 ― 30

Figure 5.5: Schematic showing the change in fluid saturation in a core sample from the initial
reservoir condition until it reaches the surface. The numbers displayed are arbitrary numbers
that are only used to explain the concept.

5.3 Summary

Fluid saturation is a percentage that indicates how much fluid the pore space
inside a rock contains, which is defined as the volume of fluid in a rock divided
by its pore volume. In reservoir rocks, the fluids are usually hydrocarbons or
water. Some extraction methods used to measure fluid saturation include retort
distillation and the Dean-Stark method. In both of these methods, the fluid is
extracted from the rock sample and then measured. The Dean-Stark method
can only measure water saturation, unlike retort distillation that can measure
both oil and water saturations. Therefore, material balance analysis is used to
supplement the Dean-Stark method to calculate the oil saturation as well. Drilling
muds make it difficult to evaluate the reservoir’s saturation using extraction
methods because they interfere with the saturation of the extracted samples.
Similarly, extreme temperatures and pressures while extracting the core samples
also cause changes within them, due to which the extraction methods cannot
accurately determine the saturation within the reservoir.

A summary of fluid saturation is presented in Table 5.1.


107
Table 5.1: Definition of fluid saturation and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Fluid Si (where i The fraction of We use fluid


saturation can be water, pore volume saturation to
oil, or gas) occupied by quantify the volume
the fluid. of oil and/or gas in
the reservoirs.

End of Chapter Questions

Question 5.1

If a core sample with a pore volume of 22.5 cm3 contains an oil volume of 9.8 cm3
and has a gas saturation of 0.15, what is its water saturation?

Question 5.2

A cylindrical core has a diameter of 1 in, length of 1.8 in, and a porosity of 19.5%.
Using solvent extraction, 1.94 cm3 of water were collected from the core and the
volume of oil was found to be 1.72 cm3. Calculate water, oil, and gas saturations.

Question 5.3

The dry weight of a rock is measured to be 211.6 g, and its saturated weight is
measured to be 219.4 g. The rock has a length of 6.5 cm, a diameter of 3 cm, and
a porosity of 18.5%. Given that the rock contains only water (ρw = 1.02 g/cm3) and
oil (ρo = 0.84 g/cm3), calculate the water and oil saturations.

108
Question 5.4

Solvent extraction was used in order to measure the fluid saturation in a core.
The following data are given for the core:

- Length: 3.5 inches

- Diameter: 2 inches

- Porosity: 22.9%

- Initial weight of the core: 487.6 g

- Final dry weight of the core: 455.9 g

- Water volume measured: 14.23 cm3

- Water density: 1.04 g/cm3

- Oil density: 0.80 g/cm3

Calculate the fluid saturations.

Question 5.5

A core contains only oil and water within it. Find the water and oil saturations by
using the following data:

- Dry weight: 168.3 g

- Saturated weight: 203.2 g

- Weight of the core after water removal: 184.1 g

- Diameter: 3.62 cm

- Length: 6.92 cm

- Water density: 1.03 g/cm3

- Oil density: 0.79 g/cm3

109
Question 5.6

A Dean-Stark apparatus was used to determine the fluid saturations in a


cylindrical core of diameter 3 cm, length 5.8 cm, and porosity 15.2%. The initial
weight of the core prior to extraction was measured to be 151.6 g. The weight
of the core plug after it was extracted and dried was 146.3 g. The volume of
water recovered from the core in the experiment was 2.94 cm3. The densities of
reservoir water and oil are 1 g/cm3 and 0.86 g/cm3, respectively. Determine the
initial fluid saturations in the core.

110
111
112
Chapter 6

Electrical
Properties
As discussed in Chapter 5, there are two methods to estimate fluid saturation:
direct and indirect. One way of estimating fluid saturation indirectly in cores
or reservoirs is by understanding and using electrical properties of reservoir
rocks. Understanding the concepts of electrical properties in core samples is
very simple, since the principles of electrical circuits apply here. The concepts
rely on understanding that different materials conduct electricity differently. For
instance, iron is considered a good conductor of electricity when compared with
wood. Similarly, water is a better conductor of electricity when compared with
hydrocarbons. Electrical conductivity is a measure of the ability of a material to
transmit electricity. However, we will study resistivity, which is the inverse of
conductivity and a measure of the material’s ability to resist the flow of electricity.
Obtaining fluid saturation from electrical properties involves correlating different
resistivities by examining rocks at different experimental conditions. Archie first
found this correlation in 1942. In the following sections, we will explore Archie’s
law and its applications in finding fluid saturation in both core samples and
reservoirs.

6.1 Understanding Archie’s Law

6.1.1 Introduction to Ohm's Law

Before discussing Archie’s law, we need to understand the concepts behind basic
electrical circuits. First, we start with Ohm’s law:

∆V = Ir (6.1)
where ΔV is the potential difference or voltage [V], I is the electric current [A], and
r is the resistance opposing the flow of electrical current [Ω].

On the other hand, resistivity is defined as:

rA
R= (6.2)
L
where R is the resistivity [Ω.m], A is the cross-sectional area [m2], and L is the
length [m].

We need to distinguish between resistance and resistivity. Resistance is an


extensive property, which means it is size dependent; this means that differently
sized samples will have different values of resistance. Another example of an
extensive property is volume. On the other hand, resistivity is an intensive

113
property; this means that the property is independent of size. Density, for
instance, is an intensive property, since a droplet of water and a giant pool of
water both have a density of 1000 kg/m3 and change in size would not change
the value of density.

Flow of Electrical Charges

Before discussing Archie’s law, let us understand the flow of electrical charges.
Let us substitute Equation 6.2 in Equation 6.1 to obtain:

IRL
∆V = (6.3)
A
Conductivity is the inverse of resistivity and is expressed as:
1
C= (6.4)
R
where C is the conductivity [1/Ω.m or Siemens per meter, S/m].

Now, we substitute Equations 6.4 in Equation 6.3 and rearrange the equation
to obtain:
CA
I= ∆V (6.5)
L
Note that this equation is similar to Darcy’s law for linear incompressible fluid flow
in porous media (Equation 4.8). The electrical current in this case is analogous to
the flow rate; the potential difference is similar to the pressure difference, where
the flow occurs from high to low voltage, and the conductivity resembles the
permeability over viscosity. The flow equations generally have the same format
for all flow types.

6.1.2 Formation Factor


The first parameter in Archie’s law is the formation factor, F. Archie’s law is
based on empirical (experimental) correlations. The formation factor is a ratio of
water resistivity (Rw) to the resistivity of a core sample fully saturated with water
(Ro), as shown in Figure 6.1. Mathematically, the formation factor is expressed as:
Ro
F = (6.6)
Rw
where F is the formation factor [dimensionless], Rw is the resistivity of water
[Ω.m], and Ro is the resistivity of the core when it is 100% saturated with water
(Sw=1) [Ω.m].

Based on experimental observations, the formation factor can also be expressed


as:
Ro a
F = = aφ−m = m (6.7)
Rw φ

114
where a is an empirical constant [dimensionless] and usually equals 1, ф is the
porosity [dimensionless], and m is a cementation factor or exponent and usually
equals 2 (the higher the cementation in a rock, the higher the m value).

Consider two rock types with different degrees of cementation. We take plug
samples of each rock type (of different porosities), and saturate them all with
water or brine and measure their resistivity (we will discuss the details of
measuring the resistivity in the following sections). A log-log plot of formation
factor as a function of porosity can be generated (Figure 6.2). As we can see,
both trend lines converge at 1, which indicates that the empirical correlation
is logical. This is because, as the porosity approaches 1, the Ro approaches Rw,
meaning that the system acts as a tank of water. Based on this correlation, we
can use the logarithmic or natural logarithmic rules to find the parameters. The
correlation that will be used later on is:

log F = log a − m log φ (6.8)

This equation resembles the straight line equation, where –m is the slope of the

line. We can solve the equation at any point with m to find a. Bear in mind that m
is positive and the negative sign is to account for the negative slope.

I V I V
+ +

Rw Ro
A A

L L
(a) (b)

Figure 6.1: Schematic showing a box with a length (L) and cross-sectional area (A) with (a) a box
that is fully filled with water (the resistivity of this box is Rw) and (b) a box that resembles a rock
with a certain porosity which is filled entirely with water (Sw =1) (the resistivity of this box is Ro).

115
1000 3

100 2

F Rock Type 2 Rock Type 1 log F Rock Type 2 Rock Type 1

10 1

Power trend line Linear trend line


F = aф-m log F = -m log ф + log a

1 0
0.1 1.0 -1 0
ф log ф
(a) (b)

Figure 6.2: Schematic showing the formation factor as a function of porosity for two rock types
by plotting (a) F and ф on a log-log scale and (b) log F and log ф on a linear scale. The cementation
exponent m can be found using a power trend line and a linear trend line for a and b, respectively.
Each data point for each rock type represents a measurement conducted on one core sample.

Example 6.1
A formation water’s resistivity is 0.7 Ω.m, and the formation rock that is
100% saturated with this water has a resistivity of 20.4 Ω.m. Given that a = 1
and m = 2, determine the porosity of this formation rock.

Solution
From the question, the values of Ro and Rw are given:

Ro = 20.4 Ω.m, Rw = 0.7 Ω.m

Hence, the formation factor F can be found using Equation 6.6:



F =
Ro
Rw
=
20.4
0.7
= 29.14
Now, we can rearrange Equation 6.7 to find porosity:


 a 1/m  1 1/2

φ=
F
=
29.14
= 0.185

116
6.1.3 Resistivity Index

Archie compared Ro to Rt, which is the true resistivity of a core that may contain
water and hydrocarbons at different saturations, as shown in Figure 6.3,
unlike Ro which has 100% water saturation. The ratio of Rt over Ro is known as
the resistivity index, Ir, or the saturation index/equation. Mathematically, the
resistivity index is expressed as:

Rt
Ir = (6.9)
Ro

where Ir is the resistivity index [dimensionless], Rt is the true resistivity of a core


sample [Ω.m], and Ro is the resistivity of the core when it is 100% filled with water
[Ω.m].

Based on empirical (experimental) observation, the resistivity index is also equal


to:
Rt 1
Ir = −n
= Sw = n (6.10)
Ro Sw
where Sw is the water saturation [dimensionless] and n is the saturation exponent
[dimensionless]; the saturation exponent is usually 2.

In order to analyze the resistivity index, a graph of resistivity index as a function


of water saturation on a log-log scale is obtained (Figure 6.4). The data points
for one curve represent one core at different water saturations, which can be
achieved in the laboratory. Unlike the formation factor where several plugs are
used, only one core sample is used for several measurements for resistivity
index. We will go through this in more detail in the following sections. Now, we
can apply the logarithmic rules to Equation 6.10 to obtain:

log Ir = −n log Sw (6.11)

where –n is the slope of the line. Again, n should be positive as the negative sign
in the equation is to neutralize the decreasing slope.

I V I V
+ +

Ro Rt
A A

L L
(a) (b)

Figure 6.3: Schematic showing a box with a length (L) and cross-sectional area (A) with (a) a
box that resembles a rock of a certain porosity which is filled entirely with water (Sw =1) (the
resistivity of this box is Ro) and (b) the same rock with both water and hydrocarbon saturations
(the resistivity of this box is Rt).

117
1000 3

100 2

Ir Rock Type 2 Rock Type 1 log Ir Rock Type 2 Rock Type 1

10 1

Power trend line Linear trend line


Ir = Sw-n log Ir = -n log Sw

1 0
0.1 1.0 -1 0
Sw log Sw

(a) (b)

Figure 6.4: Schematic showing the resistivity index as a function of water saturation for two
rock types by plotting (a) Ir and Sw on a log-log scale and (b) log Ir and log Sw on a linear scale. The
saturation exponent n can be found using a power trend line and a linear trend line for a and b,
respectively.

6.1.4 Archie’s Equation

Since we have covered the two main components of Archie’s equation, formation
factor and resistivity index, we can now discuss the final form of Archie’s law.
Bear in mind that the use of formation factor and resistivity index on their own
is not significant; however, combining them together will enable us to find the
water saturation from resistivity measurements (Figure 6.5). Note that the
term Ro is common in both Equations 6.7 and 6.10; thus we can rearrange
Equation 6.7 to obtain:

Ro = aφ−m Rw (6.12)

Note that Ro occurs in a special scenario, and thus eliminating it makes the
equation more representative. This is achieved by substituting the above
equation in Equation 6.10 and rearranging the equation to obtain:

aRw
Rt = aφ−m Rw Sw
−n
= (6.13)
φ m Sw
n

Then, we make Sw as the subject:

aRw
n
Sw = (6.14)
φ m Rt

118
Finally, we can take the nth root of both sides for this equation to obtain:


aRw
Sw = n
(6.15)
φ m Rt

This is the final form of Archie’s equation. Table 6.1 shows a description of all the
parameters in Archie’s equation.

Rt Ro Rw
ф = 20% ф = 20% ф = 100%
Sw = 20% Sw = 100% Sw = 100%

Resistivity Index Formation Factor

Resistivity

Figure 6.5: Schematic showing the two components of Archie’s law: formation factor and
resistivity index. Formation factor is a relationship between Ro and Rw, while resistivity index
is a relationship between Rt and Ro. Ro is the common parameter between the two components.

Example 6.2
A cylindrical core, of 2.5 cm in diameter and 5 cm in length, has a porosity
of 23%. The retort distillation method was used in order to find the fluid
saturation inside the core. Using this method, 3.2 cm3 of water and 1.9 cm3
of oil were extracted from the core.

a) What are the fluid saturations in the core?

b) If the formation factor is given by F = 1.09ф-2, the resistivity


index is given by Ir = Sw-2, and the water resistivity is 7.5 Ω.cm,
calculate the true resistivity of the core which we would expect
to have it measured before fluid extraction.

Solution
a) First, the bulk volume of this core needs to be calculated. Since this is a
cylinder:
 2
× 5 = 24.54 cm3
2.5
Vb = πr2 L = π ×
2

119
We now find the pore volume:

Vp = φVb = 0.23 × 24.54 = 5.64 cm3


The water saturation can be found using Equation 5.1:



Sw =
Vw
Vp
=
3.2
5.64
= 0.567

The oil saturation can be found using Equation 5.2:

So =
Vo
=
1.9
= 0.337
Vp 5.64

The gas saturation can then be found using Equation 5.4:


S w + So + S g = 1

Sg = 1 − Sw − So
Sg = 1 − 0.567 − 0.337 = 0.096
b) From the question, the values of a, m, and n are known:

a = 1.09, m = 2, n = 2

Therefore, Equation 6.13 can be used to find the true resistivity:

R = aRw = 1.09 × 7.5 = 480.7 Ω.cm


t
φ m Sw
n 0.232 × 0.5672

Archie’s equation can be mainly used to translate resistivity measurements from


wireline logging tools to water saturation; however, it cannot be used directly as
all the parameters of Archie’s equation need to be measured in the laboratory
beforehand.

120
Table 6.1: Summary of all the parameters used in Archie’s equation. The symbol [–] indicates
that the parameter is dimensionless.

Parameter Definition Unit

Sw Water saturation -

ф Porosity -

Rw Water resistivity Ω.m

Ro Rock resistivity when the rock is 100% Ω.m


saturated with water

Rt True resistivity (including hydrocarbons in Ω.m


the rock)

m Cementation exponent (usually 2) -

n Saturation exponent (usually 2) -

a Empirical constant (usually 1) -

Example 6.3
The following data is given for a core:

Diameter = 2.35 cm Length = 4.1 cm


Water resistivity = 48 Ω.cm Current = 0.02 A

Water Saturation, Sw [%] Voltage Across Core [V]

100 7.83

86 10.64

76 13.71

63 19.95

55 26.86

49 33.98

Calculate the formation factor (F) and the saturation exponent (n) of
the core.

121
Solution

The cross-sectional area of the core is:


 2
= 4.34 cm2
2.35
A = πr2 = π ×
2

When the water saturation, Sw, is 100%, the voltage is 7.83 V.

By rearranging Equation 6.3, the resistivity Ro can be found when Sw is 100%:


IRo L
∆V =
A

= 414.4 Ω.cm
∆V A 7.83 × 4.34
Ro = =
IL 0.02 × 4.1

The formation factor F can be found using Equation 6.6:


= 8.63
Ro 414.4
F = =
Rw 48

Using a rearranged Equation 6.3, the true resistivity Rt can be found for
each Sw value:
∆V A
Rt =
IL

Then, the resistivity index can be measured using Equation 6.10. Finally,
using the given data, if we plot log Sw against log Ir, the negative of the slope
of the line of best fit will give the saturation exponent n (Equation 6.11):

log Ir = −n log Sw

122
The data points are listed in the table below:

Water Saturation, Sw [-] log Sw True resistivity, Rt Ir log Ir


[Ω.cm]

1 0 414.4 1 0

0.86 -0.0655 563.1 1.359 0.133

0.76 -0.1192 725.6 1.751 0.243

0.63 -0.2007 1055.9 2.548 0.406

0.55 -0.2596 1421.6 3.430 0.535

0.49 -0.3098 1798.5 4.340 0.637

The graph of log Sw against log Ir is plotted below:


0.7

0.6 y= -2.0513x

0.5

0.4
log Ir

0.3

0.2

0.1

0
-0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 0
log Sw

The line of best fit has been found using MS Excel, and its slope is -2.0513.
Therefore, n ≈ 2.05. Note that this slope can also be calculated manually by
taking any two data points on this plot.

123
Example 6.4
The following data is given for a core:

Porosity, ф [%] Formation Factor, F

16.5 45

18.1 34

19.4 26

21.0 21

22.3 18

23.2 16

Calculate the cementation factor (m) and empirical constant (a) of the core.

Solution
This question is similar to the previous example. Equation 6.8 gives us:

log F = log a − m log φ

Therefore, using the given data, if we plot log ф against log F, the negative of
the slope will give the cementation factor m, and the y-intercept will give the
log of the empirical constant a.

The data points are listed in the table below:

Porosity, ф [-] Formation Factor, F log ф log F

0.165 45 -0.783 1.653

0.181 34 -0.742 1.531

0.194 27 -0.712 1.431

0.21 21 -0.678 1.322

0.223 18 -0.652 1.255

0.232 16 -0.635 1.204

124
The graph of log ф against log F is plotted below:

1.7

1.6
y= -3.0501x-0.736

1.5
log F

1.4

1.3

1.2

1.1
-0.80 -0.78 -0.76 -0.74 -0.72 -0.70 -0.68 -0.66 -0.64 -0.62 -0.60

log ф

The line of best fit has been found using MS Excel, and its slope is -3.0501.
Therefore, m ≈ 3.05.

Furthermore, the intercept of this line has been found to be -0.736. Since
the intercept is log a:
log a = −0.736

a = 10−0.736 = 0.184

6.2 Factors Affecting Resistivity of Reservoir Rocks

There are several factors that affect the resistivity of core samples, but we will
focus only on four of these factors:

1) The type of fluid in the pore spaces. For instance, water is a better
conductor of electricity than oil, and oil is a better conductor than gas.
If you examine water by itself, brine is more conductive than fresh
water. This is because brine has ions such as Na+, Cl-, and K+ that
provide better conductance of electricity. Nevertheless, resistivity is the
inverse of conductivity, which means that among all fluids discussed,
gas has the highest resistivity, while brine has the lowest.

2) Porosity of the formation. We can see from Archie’s equation that


porosity is a factor affecting resistivity. The higher the porosity, the
lower the resistivity, as the rock on its own has a higher resistivity than
the fluids present in it. By increasing the porosity, we subject the core to
contain more fluids, which leads to a reduction in the core’s resistivity.

125
3) Presence of clays in the rock. Clays are conductive to electricity due
to high water/brine content and thus their presence in the rock increases
the electrical conductivity, thus reducing the resistivity (Figure 6.6).

4) Degree of cementation. Rocks with higher degree of cementation


tend to have lower porosity, which leads to a higher resistivity.

100

Without conductive solids

Ir 10

With conductive solids

1
1 10 100

Sw [%]

Figure 6.6: Schematic showing the effect of clays on the resistivity index versus water saturation
relationship.

6.3 Measuring Electrical Properties of Reservoir Rocks

Measuring the electrical properties can be divided into two parts: calculating the
formation factor and the resistivity index.

6.3.1 Measuring the Formation Factor

For this test, several core samples are required, especially if we are dealing with
heterogeneous rocks such as carbonates, which have a wider porosity range
compared to sandstones. Moreover, formation factor is a function of porosity

126
and thus the variation of each rock type is a function of heterogeneity (diverse
properties) of the rock sample. Generally speaking, sandstones are more
homogeneous compared to carbonates and hence the spectrum of data will be
smaller. In order to measure the formation factor, we first take the dimensions
of the core sample and saturate it with water. This can be done either by
submerging the core sample in brine, or by using a core holder similar to what
was discussed in Chapter 4. Then, we place a core in a system similar to Figure 6.7
and apply low current from the AC source, and read the voltage from the
voltmeter. Some systems use output resistance directly from where resistivity is
measured. In the system displayed, a resistor with a known resistance is placed
for the purpose of calibration. In addition, low current is preferred to avoid the
core being heated up. From this measurement, we obtain the resistance of the
core and thus we may use Equation 6.2 to obtain the resistivity and Ro of the core
sample. We also need to measure the resistivity of water using a liquid resistivity
meter, which should be a constant value. We can then find the formation factor
for this particular rock. The same procedure is repeated for all the core samples;
hence, several data points are generated and a graph similar to that in Figure 6.2
can be plotted.

6.3.2 Measuring the Resistivity Index

The procedure is similar to that for the formation factor. However, the main
difference is that instead of using all the core samples, we use only one core to
generate an entire curve. First, we start with the Ro value obtained in the previous
test. We then decrease the saturation (desaturation) by an interval, and measure
the resistance and resistivity periodically. We can desaturate the core using
a core holder, injecting oil to replace the water. The core is extracted and the
resistivity is measured; this resistivity is equivalent to Rt. After generating several
data points, a graph similar to that in Figure 6.4 can be plotted.

AC Source

Resistor

Core

Voltmeter

Figure 6.7: Schematic showing the experimental set-up to measure the electrical properties of
core samples.

127
6.4 Applications of Electrical Properties of Reservoir Rocks

Electrical properties can help us find the water saturation not only in core samples
but also in reservoirs by using resistivity wireline logging tools. Using resistivity in
core samples can be coupled with different concepts such as capillary pressure
(Chapter 8) and relative permeability (Chapter 9).

The main application of electrical properties is to calibrate the resistivity log in


wireline logging (Figure 6.8). We divide this log into three tracks: gamma ray,
resistivity, and porosity tracks. The gamma ray track would tell us whether the
formation is a shale or reservoir rock. A high gamma ray response indicates the
presence of shale, while a low gamma ray response indicates reservoir rock.
From this track, we can identify the thickness of the reservoir formation, which in
this case is sand. The reservoir rock is where the hydrocarbons reside.

If we now go to the second track, we can identify the parts of the sand formation
that contain hydrocarbons, as well as the parts that hold formation water or
brine. The resistivity difference between gas and oil is not that significant for the
resistivity log to spot it. We will talk about how we can distinguish between the
phases when we discuss track three. However, before we do that, it is important
to note that resistivity logs can be classified as shallow or deep. Shallow resistivity
logs have a small depth of investigation, and that limits them to only measure
the resistivity in the area where the drilling mud has invaded, as we discussed
in Chapter 5. Measuring the resistivity in that area is not representative of
the reservoir’s true saturations. Deep resistivity logs, on the other hand, have
a deeper range of investigation that can go beyond the area affected by the
drilling mud. This type of log measures the resistivity of the unaffected area
which represents the reservoir’s original conditions. The resistivity log in Figure
6.8 represents the deep resistivity log. The deep resistivity log can distinguish the
hydrocarbons from brine. As we can see in the figure, hydrocarbons have high
resistivity when compared to brine.

The porosity logs in the third track can assist us in identifying the type of
hydrocarbons in the reservoir. The first indication of gas being present in the
reservoir is the crossover of the neutron and density logs. In such a case, the
density log gives low bulk density readings because the density of gas is low
compared to that of oil and water. In addition, the neutron log also gives low
readings because gas has lower number of hydrogen atoms when compared
to oil. This effect decreases when we go to the oil layer as shown in Figure 6.8.
There is also a minor reduction in crossover between the oil and water layers.
Nonetheless, it is difficult to distinguish that and the best way to distinguish
between the layers is to refer back to the resistivity log in the second track.
Relying on one track is not conclusive to provide a clear saturation distribution
of the reservoir, and therefore validating the three tracks is essential to better
understand the reservoir.

128
In order to find the water saturation in the reservoir, a fourth track can be
created. However, before we do that, core samples need to be extracted from
the reservoir to estimate the Archie parameter and to perform calibration. In
addition, a sample of the formation water is required to measure the resistivity
of water. After obtaining all the parameters, one of the porosity logs, such as the
neutron log, can be used in the computation for Archie’s equation in the fourth
track. The porosity and true resistivity will vary with depth, but other parameters
will stay constant for a specific rock type.

NPHI
0.45 -0.15

Gamma Ray Depth, Resistivity RHOB


0 gAPI 150 ft 0.2 Ω.m 20 1.90 g/cc 2.90
6,000

Shale

Gas
6,100
Hydrocarbon

Sand Oil

6,200
Brine Brine

Shale

6,300

Figure 6.8: Schematic showing wireline logging tools with three tracks: (left) gamma ray, (middle)
resistivity log, and (right) porosity logs.

6.5 Summary

Fluid saturation in rock samples can be estimated by analyzing their electrical


properties. More specifically, the resistivity measurements of core samples are
generally correlated with their fluid saturation using Archie’s equation to find the
water saturation from resistivity measurements. The two main components of
Archie’s equation are formation factor (F) and resistivity index (Ir). The formation
factor is the ratio of water resistivity (Rw) to the resistivity of a core sample fully
saturated with water (Ro), whereas the resistivity index is the ratio of the true
resistivity of the core (Rt) to the resistivity of a core sample fully saturated with
water (Ro). Archie's parameters (a, m, and n) are measured in the laboratory

129
from extracted core samples. Some of the factors that affect the resistivity of
core samples include the type of fluid in the pore spaces, the porosity of the
formation, the presence of clays in the rock, and the degree of cementation.
The formation factor is obtained by measuring the resistivity of several rock
samples as well as that of water, whereas the resistivity index is obtained by
measuring the resistivity of only one sample. Electrical properties have several
applications and are not only limited to obtaining fluid saturation in rocks.
The resistivity of core samples can also be coupled with capillary pressure and
relative permeability within rocks, which will be discussed in Chapters 8 and 9.
Electrical properties are also used in wireline logging to determine the thickness
of reservoir formations and the type of hydrocarbon in the reservoir, as well as to
distinguish hydrocarbons from brine. Table 6.2 summarizes electrical properties
in reservoir rocks.

Table 6.2: Definition of electrical properties and their importance to the


petroleum industry.

Parameter Symbol Definition Importance

Electrical F, Ir The rock properties We use electrical


Properties that impede the flow properties obtained
(formation of electric current from the laboratory
factor and through a rock. to calibrate
resistivity Electrical properties resistivity logs
index) are a function of and find the fluid
porosity and fluids saturation in the
in the pore spaces. reservoir.

130
End of Chapter Questions

Question 6.1

What is the true resistivity of a rock whose water saturation is 0.45 and whose
resistivity when 100% saturated with water is 32 Ω.m? Assume a saturation
exponent n of 2.

Question 6.2

A rock sample has a formation factor of 25, and the water within it has a resistivity
of 1.2 ohm-m. Calculate:

a) The resistivity of the rock when it is 100% saturated with water

b) The porosity of the rock. Assume a = 1.1 and m = 2.5.

Question 6.3

A rock of unknown brine saturation was used in a flooding experiment where


brine resistivity was measured at 7 Ω.cm. The rock has a 1.2 in diameter and a 5.4
in length, with a porosity of 9.4%. A resistance of 23286 Ω was measured across
the rock. The following equations describe the system:

- F = 0.93ф-2

- Ir = Sw-1.9

Calculate the brine saturation in the rock.

Question 6.4

a) A core sample with a diameter of 1.2 in and a length of 1.8 in is saturated


with brine solution of 0.26 Ω.m resistivity. The resistance of the core
when fully saturated with brine is measured to be 1549.3 Ω. Calculate
the formation factor F for this core.

b) Assuming a = 1 and ф = 17.6%, what is the value of the cementation


exponent m for this core?

131
Question 6.5

A cylindrical core has a 2 in diameter, 3.5 in length, and a porosity of 18.1%. Using
retort distillation, 15.3 cm3 of water and 12.4 cm3 of oil were extracted from the
core. If the formation factor of this rock is given by F = 1.26ф-2.1, the resistivity
index is given by Ir = Sw-1.9, and the water resistivity is 8.4 Ω.cm, calculate the true
resistivity of the core which we would expected to have before fluid extraction.

Question 6.6

In an attempt to derive Archie’s equation parameters, three reservoir rock


samples were identified and tested. Length, diameter, and dry weight were
measured. The rocks were subjected to strong vacuum and then saturated
with brine of 1.03 g/cm3 and 7.3 Ω.cm resistivity. The saturated weight was then
measured. The electrical resistance was measured using a voltmeter with a
constant current. Using the data of the three rocks given below, calculate Archie’s
parameters (a and m).

Rock L [in] D [in] Wdry [g] Wsat [g] r [Ω]

R1 2.3 1.02 83.4 86.1 193

R2 1.7 1.01 72.5 75.4 76

R3 2.9 1.01 85.3 93.4 56

Question 6.7

a) Several core plugs (diameter 1 in and length 1.5 in) were taken
from a given reservoir. Each was cleaned and its porosity was
measured using a given gas expansion porosimeter. Then, all were saturated
with brine having a resistivity of 7.5 Ω.cm. After that, each core plug was
placed in a resistivity apparatus to measure the voltage drop under 0.01 A
current. The porosity and voltage drop measured are listed below:

Sample Number Porosity, ф [%] Voltage Across Core [V]

1 17.8 1.60

2 18.8 1.46

3 16.5 1.96

4 22.0 1.09

5 15.5 2.14

6 14.5 2.41

Calculate F and determine the parameters a and m.

132
b) The laboratory procedure was continued using Sample No. 1. The
brine saturation was reduced step-by-step by displacing the core plug with oil,
and then the voltage drop was measured.

Water Saturation, Sw [%] Voltage Across Core [V]

100 1.60

70 3.06

52 4.86

43 7.30

35 10.63

Calculate the resistivity, Rt, as a function of Sw, and determine n.

c) Calculate the water saturation when the porosity is 0.17 and the true resistivity
is 537 Ω.cm.

133
134
Chapter 7

Wettability
Wettability is the preference of a solid to be in contact with one fluid over another
in a system of two or more immiscible fluids. The concept of wettability can be
observed in our daily lives: for instance, when a droplet of water spreads on a
tissue paper, it indicates that the tissue paper prefers to be in contact with water
or the tissue paper “likes” the water (hydrophilic). In this case, we can say that
the tissue paper is water-wet, which means that it prefers to stay in contact with
water rather than any other fluid. Similarly, the fabric used in making umbrellas
“does not like” to be contacted by water, which explains why rain droplets slide
easily off of umbrellas (hydrophobic). In this case, the fabric is known to be non-
water-wet or oil-wet. In the same way, reservoir rocks tend to have a preference
to be in contact with either oil or water. Knowing the wettability is very important
in understanding flow behavior when we are injecting water to displace oil. For
example, if we are injecting water (waterflooding) in a water-wet rock (rock that
likes water), the flow will be different from the case of an oil-wet rock. This is
due to the low mobility of water in a water-wet rock as water wants to stick to
the surface while the oil will be expelled out easily. Therefore, wettability can be
considered as a flow or transport property that affects the flow in porous media.
We will discuss this in more detail in the following sections.

7.1 Understanding Wettability

7.1.1 Surface and Interfacial Tension

Figure 7.1 shows a beaker filled with water up to a certain point and then filled
with air: point (a) in Figure 7.1 is a water molecule that is balanced; however,
point (b) in the same figure is imbalanced as the water molecule is not fully
connected to water. This causes a tension, which is commonly known as surface
tension (ST). Surface tension occurs between water and gas or between a solid
and fluid (liquid or gas). Interfacial tension (IFT), on the other hand, takes place
between two liquids, and the lower the interfacial tension between two fluids,
the closer the fluids will become to being miscible (miscibility is the ability of two
fluids to mix together). In either case, it is defined as the energy per unit area or
the force per unit length [N/m], which can be expressed as [dyne/cm, where 1
dyne = 10-5 N] with the symbol σ. Figure 7.2 shows a clear interface between two
fluids (high interfacial tension) and the case when there is no clear interface (low
interfacial tension). Interfacial tension is closely connected to wettability, which
we will explore in the following sections. It is a function of the fluid pairs as well
as the temperatures and pressures that the fluid pairs are subjected to. Table
7.1 shows typical surface/interfacial tension values for specific fluid pairs.

135
(b)

(a)

Figure 7.1: Schematic showing a beaker filled with water to a certain point, and then filled
with air. Point (a) in the schematic shows a balanced water molecule while point (b) shows an
imbalanced water molecule.

Oil

Water

Lower interfacial tension


(a) (b)

Figure 7.2: Schematic showing (a) a clear interface between oil and water which indicates higher
interfacial tension when compared to (b) a hypothetical case with no visible interface which
indicates that oil and water are partially miscible due to the low interfacial tension (since oil and
water are immiscible).

Table 7.1: Typical ST/IFT values for specific fluid pairs at ambient conditions.

System σ [mN/m]

Air/mercury 480

Gas/oil 24

Gas/brine 72

Oil/brine 32

136
7.1.2 Adhesion Tension

When we deal with forces in molecules, we can talk about two types of forces:
cohesive and adhesive. Cohesive forces are forces of attraction between similar
molecules while adhesive forces are forces of attraction between different
molecules (Figure 7.3).

We have cohesive forces between water molecules that prefer to stay together
and adhesive forces that prefer to spread out on solid surfaces. This system,
for instance, is water-wet as the water droplet prefers to spread on the surface
over staying in contact with the water molecules, which means that the adhesive
forces are stronger than the cohesive forces.

Moreover, if we want to analyze a droplet of water on a solid surface that is static


(Figure 7.4), then we can analyze the forces acting on the horizontal plane of
point A of that system:

Σfx = σso − σsw + σwo cos θ = 0 (7.1)

where Σfx are the forces in the horizontal plane, σsw is the surface tension between
the solid and water [N/m], σso is the surface tension between the solid and oil
[N/m], σwo is the interfacial tension between water and oil [N/m], and θ is the
contact angle between the water droplet and the surface.

Adhesion tension can be defined as the difference between two solid-fluid


surface tensions:

AT = σsw − σso = σwo cos θ (7.2)

By rearranging the parameters to find the contact angle, the equation becomes:

σsw − σso AT
cos θ = = (7.3)
σwo σwo

A positive adhesion tension indicates that the denser phase (in this case water)
wets the surface and vice versa for a negative adhesion tension. On the other
hand, an adhesion tension of zero indicates that the system has neutral affinity
towards both fluids.

137
Gas

θ Water

Solid

Cohesive force
Adhesion force

Figure 7.3: Schematic showing a droplet of water surrounded by gas on a solid surface. The blue
arrows depict cohesive forces between the water molecules, and the red arrows depict adhesive
forces between the water molecules and the solid surface. The arrows at the bottom of the
figure indicate that cohesive and adhesive forces are acting in opposite direction. The contact
angle θ is the angle between the liquid and the surface.

+y

σwo cos θ

-x A +x
-x +x
σso σsw

Oil Water Solid

Figure 7.4: Schematic showing a droplet of water surrounded by oil on a solid surface in static
equilibrium, where the summation of the surface and interfacial tensions on the horizontal
plane should be equal to zero. s, w, and o denote solid, water, and oil, respectively.

7.2 Classification of Wettability

In an oil-water reservoir, wettability can be divided into four categories:

1) Water-wet, where the rock surface prefers to be coated with water


and thus the rock has a high affinity towards water, allowing water to
spread on the surface (Figure 7.5a). This means that the contact angle
will be less than 90° as the water spreads on the surface.

138
2) Intermediate-wet or neutral-wet, where the surface has an almost
equal tendency to be coated by one of the fluids (either oil or water)
(Figure 7.5b). This means that the contact angle is around 90° as the
surface has equal affinity towards both oil and water.

3) Oil-wet, where the rock prefers to be in contact with oil, opposite to


the water-wet case (Figure 7.5c). In this case, the contact angle will be
greater than 90° as the surface prefers to be in contact with oil over
water. The cohesive forces between the water molecules are stronger than
the adhesive forces between the water molecules and the surface,
and thus water droplets will stick together to form a sphere. Every
system undergoes an energy minimization phenomenon, in which the
system tends to go towards the lowest energy state and hence the droplet
takes the shape of a sphere as spheres have the smallest surface area.

4) Mixed-wet, where parts of the rock prefer to be in contact with oil and
the other parts prefer to be in contact with water (Figure 7.5d). Mixed-
wettability can also be referred to as fractional wettability and the
contact angle will vary depending on the region of the rock.

Table 7.2 summarizes the expected contact angle values depending on the
wettability state.

Clean reservoir rocks or minerals (quartz [sandstone] or calcite [carbonate])


tend to be water-wet and coated with water layers. Once these minerals stay in
contact with crude oil for a sufficient time due to oil migration, the heavy polar
components of the crude oil (resins and asphaltene) precipitate on the surface,
forming an oil layer that makes the rock more oil-wet. Similarly, when our skin
is coated with oil or paint, it becomes difficult to be displaced by water as the
surface becomes oil-wet. In this case, the only way to remove the oil from the
skin is to apply soap which will lower the surface tension between the oil and the
skin and will become miscible with the oil which will then make it easy for the
water to displace the oil. This application of using soap is similar to the injection
of surfactant (like soap) in the reservoir as part of the Enhanced Oil Recovery
(EOR) process to extract oil in the reservoir by lowering the surface/interfacial
tension.

θ < 90º θ = 90º θ > 90º

Water-Wet Intermediate-Wet Oil-Wet Mixed-Wet


(a) (b) (c) (d)

Figure 7.5: Schematic showing contact angle measurement on a solid surface for (a) water-wet,
(b) intermediate-wet, (c) oil-wet, and (d) mixed-wet cases.

139
Table 7.2: Wettability classification based on the
contact angle of the water droplet on a rock surface.

Wettability State θ [°]

Water-wet <90

Intermediate-wet ≈90

Oil-wet >90

7.3 Flow Sequence/cycle

The flow sequence in porous media can be classified into two main types:
drainage and imbibition. Drainage means the decrease in the wetting phase
while imbibition is the increase in the wetting phase. For example, the flow of
water into a water-wet tissue paper is an imbibition process as the wetting phase
(water) increases in the medium. In the reservoir, when the oil migrates from the
source rock to the reservoir, assuming that a water-wet reservoir, the process
is termed as drainage; when water is injected in the same reservoir, assuming
it maintains its initial water-wet state, to displace oil, the process is termed as
imbibition. Different properties will change depending on the flow sequence,
these are: capillary pressure (Chapter 8), relative permeability (Chapter 9), and
contact angle.

7.4 Measuring Wettability

Wettability can be measured directly or indirectly. Direct measurements include


contact angle and Amott index. Indirect measurements include imbibition
capillary pressure (Chapter 8) and waterflood relative permeability (Chapter 9).
The following section will discuss the direct measurements while the indirect
measurements will be discussed in their respective chapters.

7.4.1 Contact Angle

The apparatus to measure the contact angle consists of a high resolution camera
attached to a computer (Figure 7.6). The camera is set to be in a linear position
to the sample which is to be measured. Before taking the measurements, the
camera needs to be calibrated for focus and set in a central position using a
calibration tool. Once calibrated, a fluid droplet is placed on the intended
surface. To process the data, a software used for this experiment will identify
the contact angle based on the interface of the droplet. If the interface is not
detected by the software, then a manual identification of the interface is required.
Table 7.3 shows expected contact angle values for specific fluid pairs on a clean
glass surface.

140
Liquid dropper

Camera

Computer Stand

Figure 7.6: Schematic of the experimental apparatus to measure the contact angle. In this
system, the contact angle of a water droplet surrounded by air is measured.

Table 7.3: Expected contact angle values for specific fluid pairs
on a clean glass surface.

System θ [°]

Air/mercury 140

Gas/oil 0

Gas/brine 0

Oil/brine 30

Contact Angle Hysteresis

When dealing with multi-phase flow in porous media, the term hysteresis is
very important. Hysteresis refers to the state of a system that is dependent on
its history. Therefore, the contact angle value will vary depending on the flow
sequence which is either drainage or imbibition. The contact angle when the
wetting phase is decreasing (drainage) is known as the receding contact angle,
θR, while the contact angle when the wetting phase is increasing is known as the
advancing contact angle, θA. The static contact angle when there is no flow is
known as the intrinsic contact angle, θI.

Figure 7.7 shows contact angle hysteresis between the flooding cycles (drainage
and imbibition). The reason for this inconsistency can be attributed to the
chemical inhomogeneities and the roughness of the surface. The advancing
contact angle, θA, in imbibition (increase in water saturation) is found to be larger
than the receding contact angle, θR, in drainage (decrease in water saturation).
The intrinsic contact angle, θI, is measured at rest on a smooth surface and its
value lies between the advancing and receding contact angles.

141
Needle/Syringe
Needle/Syringe

θθAdvAdv
θθRecRec

(a) (b)

Figure 7.7: Contact angle hysteresis, depending on the flow direction, on a solid surface during (a)
imbibition with the advancing contact angle (denser phase advancing) and (b) drainage with the
receding contact angle (denser phase receding).

7.4.2 Amott Index

The Amott test quantitatively measures the wettability since the amount
spontaneously imbibed is a function of wettability. For example, when a tissue
paper is placed in a water beaker, you will note that the tissue paper imbibes
water (draws water in) even without applying any additional force. This
indicates that the tissue paper is water-wet. This can similarly be used to test
reservoir rocks. This test compares the recovered amount of oil by spontaneous
imbibition to the amount recovered by forced water injection to give the Amott
water index Iw. Spontaneous imbibition is the natural draw of the wetting phase
in the absence of any external force. One example of spontaneous imbibition
is placing sugar cubes in coffee (ignoring the dissolution process). You will see
air bubbles produced on the surface of the coffee after placing the sugar cubes
because the coffee (water based) imbibes into the sugar cubes (porous medium)
and displaces the air inside the cubes. In addition, the Amott oil index, Io, can
be estimated by comparing the amount of water recovered by spontaneous oil
imbibition to the amount recovered by forced oil injection. Figure 7.8 shows
schematics of the experimental Amott test for Iw and Io.

Sw (B) − Sw (A)
Iw = (7.4)
Sw (C) − Sw (A)

So (D) − So (C)
Io = (7.5)
So (E) − So (C)

where Iw and Io are the Amott indices for water and oil respectively, Sw(A) is the initial
water saturation present in the core sample before water spontaneous imbibition,
Sw(B) is the water saturation after spontaneous water imbibition (Figure 7.8a),
Sw(C) is the water saturation after forced water injection (Figure 7.8b), So(C) is the
oil saturation after forced water injection (Figure 7.8b), So(D) is the oil saturation
after spontaneous oil imbibition (Figure 7.8c), and So(E) is the oil saturation after
forced oil injection (Figure 7.8d). Note that this topic will be revisited in Chapter
8 when capillary pressure is discussed.
142
Amott index (IA) is the difference between the water and oil indices as shown
below:

I A = Iw − I o (7.6)

Depending on the value of Amott index, the wettability of the core sample can
be defined. The expected value for each wettability state is shown in Table 7.4.

Burette
Burette

0 0 000 00
3 30

GasGas

0 0 000 00
3 30
0 0 000 00
3 30

Water
Water
Core
Core

Oil Oil

Core
Core

(a) (b)

0 000 00 0 0 000 00
3 30 3 30

0 0 000 00 0 0 000 00
3 30 3 30
0 0 000 00
3 30

Oil Oil

(c) (d)
Figure 7.8: Schematic of the Amott experimental test with the Amott water index where we
place a rock filled with oil and water into the cell. (a) We fill the cell with brine and then the
rock will spontaneously imbibe water and displace the oil upward due to buoyancy forces
(density difference). Then, we measure the amount of oil recovered by spontaneous imbibition
and compare it to (b) the waterflood recovery to obtain the Amott water index (Equation 7.6).
Similarly, for the Amott oil index that is opposite to the Amott water index, (c) we fill the cell with
oil and the oil will displace the water for mixed-wet and oil-wet systems. Here, we show the cell
to be inverted as brine is denser than oil and will travel downward, and (d) then oil is injected
to displace the water out of the core sample to measure the amount recovered by forced oil
injection.

143
Table 7.4: Expected Amott index values for different wettability states.

Water-wet Intermediate-wet Oil-wet

Amott index (IA) 0.3 – 1.0 -0.3 – 0.3 -1.0 – -0.3

Example 7.1
A core sample initially contained Sw = 25%. The core was placed in an Amott
cell surrounded by water for a few days. After extracting the core, it was
found that the water saturation increased to 68%. When the core sample
was placed in the core holder and water was injected at a high flow rate,
the water saturation in the core after water injection was found to be 72%.
When the core sample was placed in an Amott cell surrounded by oil, very
little oil imbibition occurred and the water saturation decreased to 71%.
When the core sample was placed in the core holder and oil was injected to
displace the water, the water saturation reduced to 32%. Based on the data
collected, what is the wettability of the core sample?

Solution
First, the water and oil saturations in the problem have to be identified.
They are as follows:

Sw(A) = 0.25 Sw(B) = 0.68 Sw(C) = 0.72


So(C)= 1 - 0.72 = 0.28 So(D) = 1- 0.71 = 0.29 So(E) = 1 - 0.32 = 0.68

Then, using Equation 7.4, the Amott water index can be found:

Sw (B) − Sw (A) 0.68 − 0.25


I = = = 0.915
w Sw (C) − Sw (A) 0.72 − 0.25

Similarly, using Equation 7.5, the Amott oil index can be found:

So (D) − So (C) 0.29 − 0.28



Io = = = 0.025
So (E) − So (C) 0.68 − 0.28
Finally, the Amott index can be found using Equation 7.6:

I = I − I = 0.915 − 0.025 = 0.890


A w o

This value of the Amott index falls in the range of water-wet. Therefore, the
core sample is water-wet.

144
7.5 Applications of Wettability

The applications of wettability can be viewed in our daily lives. For instance,
rain coats, umbrellas, wool on sheep, and even ducks’ feathers are created to
be oil-wet in order to repel water and prevent it from soaking into the material
by spontaneous imbibition. On the other hand, towels and tissues, in general,
are made to be water-wet to soak up water. When dealing with reservoir rocks,
wettability helps in understanding the saturation distribution in the reservoir
at the pore-scale, as shown in Figure 7.9. Different wettability states will have
different saturation distributions at the pore-scale. In addition, the amount of oil
recovered by water injection is mainly dictated by the wettability of the rock. This
will be discussed further in Chapter 9.

Water-Wet Oil-Wet Mixed-Wet

(a) (b) (c)

Figure 7.9: Possible water and oil distribution in (a) a water-wet system where oil (green) remains
in the center of the pores and fills the large pores, (b) an oil-wet system where water (blue)
remains at the center of the pores and fills the large pores while the oil surrounds the water, and
(c) a mixed-wet system where oil has displaced water from some surfaces but is still trapped at
the center of the water-wet regions.

7.6 Summary

Wettability is the preference of one fluid to spread/adhere on a surface in the


presence of another immiscible fluid. When adhesive forces (forces of attraction
between different molecules) are stronger than cohesive forces (forces of
attraction between similar molecules), for water droplets over a surface for
example, the droplets will spread out and the surface will be termed water-wet.
Similarly, when cohesive forces are stronger than adhesive forces, in an oil-water
system, the surface will be termed as oil-wet. In an oil-water reservoir, the rocks
can be water-wet (high tendency to be coated with water), intermediate-wet (equal
tendency to be coated with water or oil), oil-wet (high tendency to be coated with
oil) or mixed-wet (both water-wet and oil-wet in different parts). Wettability can
be measured directly by measuring the contact angle between the fluid and the
surface, or by measuring the Amott index. The Amott test compares the amount
of oil recovered by spontaneous imbibition (wetting phase naturally entering
a rock without any external force) to that recovered by forced water injection
to give the water index, which can then be related to wettability. Wettability is
used to design everyday items, such as umbrellas (oil-wet) and tissues (water-

145
wet), to serve specific purposes. In the oil and gas industry, wettability helps
in understanding the saturation distribution in the reservoir and is used to
determine the amount of oil recovered by water injection.

Table 7.5 summarizes the concept of wettability.

Table 7.5: Definition of wettability and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Wettability - Tendency of a We use wettability to


fluid to spread help us understand
on a solid surface oil production when
in the presence injecting water as
of another wettability has a
immiscible fluid. primary role in water
injection performance
which will be
discussed further in
Chapter 9.

146
147
148
Chapter 8

Capillary
Pressure
Capillarity is the tendency of a liquid to rise or fall in a capillary tube due to
adhesion tension. The concept of capillarity can be seen in our daily lives
when placing a straw in a glass filled with liquid, in this case water (Figure 8.1).
Moreover, we see a rise (Figure 8.1a) and fall (Figure 8.1b), which is a function
of adhesion tension, discussed earlier in Chapter 7. If the straw is water-wet,
then the water will rise, as the water prefers to be in contact with the water-
wet surface and the adhesion tension will pull the water upward. However, if
the straw is oil-wet, the water prefers to avoid being in contact with the straw’s
surface and thus will shrink and fall below the water level.

Similarly, rocks can have throats (connections) between the pores that can act
as capillary tubes. The displacement of fluids can be in the direction of or can
oppose adhesion forces. Adhesion tension is the part of capillary pressure which
helps in determining the fluid saturation in porous media above the free water
level, which will be discussed in more detail in the following sections. Capillary
pressure is the pressure difference across the interface of two immiscible fluids
and is part of SCAL measurements.

(a) (b)

Figure 8.1: Schematic showing a glass of water with a straw: (a) the water level is rising in the
straw caused by adhesion tension, indicating a water-wet straw, and (b) the water level is falling,
indicating an oil-wet straw.

149
8.1 Capillary Rise

In this section, the derivation of the height above the free water level and capillary
pressure will be discussed from a simple capillary rise experiment conducted on
a capillary tube for water/air (Figure 8.2) and water/oil systems.

8.1.1 Water/Air System

The capillary rise in Figure 8.2 is at static equilibrium, which means that the
force pulling the water upward equals the force pulling the water downward for
a water/air system.

The force pulling the water up is the adhesion tension multiplied by the
circumference of the tube as the water sticks to the circumference of the capillary
tube:

Fup = 2πrAT = 2πrσwa cos θ (8.1)

where Fup is the force pulling the water upward in the capillary tube [N], AT is the
adhesion tension [N/m], r is the radius of the capillary tube [m], h is the height of
the water column in the capillary tube above the free water level (h =0), σwa is the
surface tension between water and air [N/m], and θ is the contact angle between
the water and surface of the capillary tube [°].

On the other hand, the force pulling the water downward is the weight of the
water column, which can be expressed by the pressure of the water at height h
multiplied by the area of the tube:

Fdown = ρw ghA = ρw ghπr2 (8.2)

where Fdown is the force pulling the water downward in the capillary tube [N], ρw is
the density of water [kg/m3], g is the acceleration due to gravity [9.81 m/s2], and
A is the area of the capillary tube [m2].

Equating Equations 8.1 and 8.2 leads to:

2πrσwa cos θ = ρw ghπr2 (8.3)

The equation can be rearranged to find h:

2πrσwa cos θ 2σwa cos θ


h= 2
= (8.4)
ρw gπr ρw gr
Capillary pressure is defined as the difference between the pressure of non-
wetting phase and the wetting phase across the interface of two immiscible
fluids:
Pc = Pnw − Pw (8.5)

150
where Pc is the capillary pressure [Pa], Pnw is the pressure of the non-wetting
phase [Pa], and Pw is the pressure of the wetting phase [Pa].

The wetting phase in this system is water while the non-wetting phase is air. The
capillary pressure in this case is the difference between the air and the water
pressures across the interface, as shown in Figure 8.2.

The air pressure acting on the interface is:

Pa, h = PA − ρa gh (8.6)

where Pa is the air pressure across the interface [Pa], PA is the atmospheric
pressure [≈101 kPa], and ρa is the density of air [kg/m3].

Hydrostatic pressure is the pressure due to gravity exerted by a fluid at


equilibrium (P =ρgh). It is a function of the vertical distance submerged while
all horizontal points should have the same pressure value. Hydrostatic pressure
increases as you go deeper underneath within a fluid. For example, the pressure
increases in seas as you go deeper from the sea level and vice versa. This explains
the negative sign next to the ρagh as the air pressure at this point is lower than
that of h=0 as the point is at higher upward elevation.

The water pressure acting on the interface is:

Pw = PA − ρw gh (8.7)

Substituting Equations 8.6 and 8.7 in Equation 8.5 gives:

Pc = PA − ρa gh − (PA − ρw gh) = (ρw − ρa )gh (8.8)

Since ρw >> ρa, ρa can be neglected and Equation 8.8 becomes:

Pc = ρw gh (8.9)

Solving for h, the equation becomes:

Pc
h= (8.10)
ρw g

Now, substituting Equation 8.10 in Equation 8.4:

Pc 2σwa cos θ
h= = (8.11)
ρw g ρw gr

Thus, Pc is equal to:

2σwa cos θ
Pc = (8.12)
r

151
This is the equation of capillary pressure for an air/water system

Pa,h = PA - ρa gh

Pc,h = ( ρw - ρa ) gh

Pw,h = PA - ρw gh

PA
h=0
PA PA

Water

Figure 8.2: Schematic showing a water-wet capillary tube placed in a beaker filled with water
surrounded by air. The water level rises in the capillary tube due to adhesion tension.

8.1.2 Water/Oil System

The capillary pressure in a water/oil system is similar to that of a water/air


system with slight differences. For example, the force lifting the water column up
is not only the adhesion tension but also the weight of the oil acting on the water
surface pushing the water up in the capillary tube, as shown in Figure 8.3. This is
neglected in the water/air system because air has a low density.

Therefore, the force pulling the water up would be:

Fup = 2πrσwo cos θ + ρo ghπr2 (8.13)

where σwo is the interfacial tension of water and oil [N/m] and ρo is the density of
oil [kg/m3].

With the same force pulling the water down as given in Equation 8.2, h becomes:

2σwo cos θ
h= (8.14)
(ρw − ρo )gr

152
The capillary pressure across in the interface will be:

Pc = PA − ρo gh − (PA − ρw gh) = (ρw − ρo )gh (8.15)


In this system, the density of oil is not neglected as it is significant when compared
to the water.

Equating Equations 8.14 and 8.15 yields:


Pc 2σwo cos θ
h= = (8.16)
(ρw − ρo )g (ρw − ρo )gr
Thus, Pc becomes:
2σwo cos θ
Pc = (8.17)
r
This equation is similar to the water/gas equation with the difference in the
interfacial tension. The capillary pressure equation can be used for any fluid pair
as long as the parameters in the equation are known. The equation assumes
that the porous medium is made up of capillary tubes, which is unrealistic for
a complex porous medium; however, this equation is commonly used in the
petroleum industry to represent reservoir rocks despite its simplicity.

The capillary rise/fall is mainly a function of adhesion tension which is thus a


function of wettability. Figure 8.4 shows a schematic of the expected water
level in capillary tubes as a function of wettability. Capillary pressure can be
used to determine the water saturation above the free water level, which will be
discussed further in the following sections.

r r

Pa Po
Pw θ Pw θ

h Air h Oil

PA PA

Water Water

(a) (b)

Figure 8.3: Schematic showing a capillary rise of water in a capillary tube surrounded by (a) air
and (b) oil.

153
θ
Oil

Oil Oil
θ Water

Water Water

(a) (b) (c)


Figure 8.4: Schematic showing a capillary tube with different wettability states placed in a water/
oil system with (a) a water-wet tube, (b) an intermediate-wet tube, and (c) an oil-wet tube.

8.2 Capillary Pressure Curves

Capillary pressure can be divided into drainage and water re-saturation, which
is commonly referred to as imbibition in the literature. The process is referred
to as water re-saturation because we will examine the behavior when water is
introduced to rocks with different wettability states that resemble the injection
of water into reservoirs to displace oil. The term water re-saturation was used,
instead of imbibition, because it is more accurate: for example, the injection of
water into an oil-wet rock is a drainage process.

8.2.1 Drainage

Drainage is the displacement of the wetting phase by the non-wetting phase.


The displacement opposes the adhesion force. When oil migrates from the
source rock into a brine saturated reservoir, this process is referred to as
primary drainage. Since oil migration is upward due to the buoyancy forces
between the oil and the brine in the reservoir, it displaces the brine downward
or laterally depending on local heterogeneity. The oil invades the larger pores
first and then progressively smaller pores as it progresses until the brine reaches
an irreducible water saturation (Swir) where the water saturation cannot be
reduced any further regardless of an increase in the applied capillary pressure.
At this stage, the water will be squeezed in the corners of the pore spaces coating
the surface of the rock in the form of layers. The irreducible water saturation is
also commonly referred to as the connate water saturation (Swc).

Oil/water transition zone (TZ) in a reservoir is a zone between the oil/water


contact (OWC) and where the water saturation is at or near its irreducible value
(Figure 8.5). In other words, the TZ is where the water saturation is changing and
not constant. The OWC is the last point in the reservoir where the water saturation
is equal to 1. In addition, the pressure at the OWC is known as the threshold
capillary pressure (Pct) which is the minimum pressure required to displace the
water. The height of the OWC will vary depending on the permeability of the
rock. Let us consider a rock to be a bundle of capillary tubes as shown in Figure
154
8.6. Higher permeability rocks can be represented with larger capillary tubes
with larger radii. The water saturation tends to be lower in larger tubes since
this will result in a lower capillary pressure, and thus lower threshold capillary
pressure. Figure 8.7 shows different capillary pressure curves as a function of
permeability. Overall, higher permeability rocks tend to have lower threshold
capillary pressure. The free water level (FWL) is the free water when capillary
pressure is equal to zero. An oil/brine reservoir may have several wettability
states which can be correlated with the drainage capillary pressure. After primary
drainage, surface-active compounds in the oil may adhere to the solid surface,
changing the wettability of the system. For instance, mixed-wet wettability
behavior tends to be located within the transition zone. The wettability in the
transition zone is mixed-wet with a higher fraction of water-wet areas near the
OWC. The wettability of rocks becomes less water-wet as we reach the top part
of the oil reservoir since most of the rock surface is contacted by oil. The oil-wet
wettability state tends to be located above the transition zone where only oil can
flow. Table 8.1 lists the important parameters in the drainage capillary pressure
curve and their definitions.

Swir

100% Oil Oil Pay Zone

Oil-Wet
Height Above FWL

Oil

Transition Zone

Water
Mixed-Wet

100% Water Saturation Pct Oil-Water Contact

100% Water Free Water Level (FWL)


Sw=1
0% Swir 100%
Water Saturation
Water-Wet

Figure 8.5: Schematic diagram of drainage capillary pressure with important labeling and
wettability distribution in the reservoir.

A B B
h
A

0 Sw 1

Figure 8.6: Schematic diagram of large and small bundles of capillary tubes that represent high
and low permeability porous media, respectively, with their counterpart capillary pressure.

155
Table 8.1: Important parameters associated with capillary pressure and their definitions.

Parameter Symbol Definition

Free water level FWL It is approximated by the level of the


OWC in an open wellbore, where Pc =0.

Oil/water OWC It is the lowest point in the reservoir


contact where no oil is present (Pc =Pct).

Oil/water TZ It is the vertical interval at which oil and


transition zone water are produced simultaneously and
water saturation is changing.

Irreducible Swir It is the minimum water saturation in a


water saturation rock forming a layer of water stuck to
the rock surface and does not move, Swc
=Swir.

Connate water Swc It is the minimum water saturation


saturation that adheres to the pores and does
not become mobile. During deposition,
the water saturation is 100%, which
decreases to 20-30% due to the
migration of oil/gas .

156
Decreasing
Permeability

Pc (+ve)
B

0 1
Sw

Figure 8.7: Schematic diagram showing the effect of permeability on capillary pressure. As the
permeability increases, the threshold capillary pressure decreases.

8.2.2 Water Re-saturation

Water re-saturation can be divided into two parts: spontaneous imbibition and
forced displacement. Imbibition is the process opposite to drainage: here, the
wetting phase displaces the non-wetting phase. To emphasize this, we often
refer to imbibition as spontaneous since Pc > 0 and water is at a lower pressure
than oil. When Pc < 0, there is forced displacement where water is at a higher
pressure than oil.

Spontaneous water imbibition is the invasion of the water into a porous medium
due to capillary forces and can only occur in water-wet and mixed-wet systems.
There are many examples of spontaneous imbibition in our daily lives. An
example is the process of a tissue paper that soaks water even without applying
any force. Another example is that when we add sugar cubes to coffee (we ignore
the dissolution process), we will see air bubbles produced on the surface of the
coffee. This is because the coffee (water based) imbibes into the sugar cube
(porous medium) and replaces the air inside the system.

In imbibition following drainage, the water fills the pore space in decreasing
order of capillary pressure. In other words, water fills the narrowest throats with
the highest capillary pressure first as the imbibition starts where the drainage
capillary ends, which is at a high positive capillary pressure. However, the
constraint is that the water cannot invade a throat unless the neighboring pore
or throat is filled with water.

157
Forced displacement or waterflooding occurs after spontaneous imbibition to
recover the oil that was not recovered by spontaneous imbibition. In forced
displacement, we need to inject water at a higher pressure than oil to displace
the oil, thus yielding a negative capillary pressure since the water pressure is now
higher than the oil pressure. The water will keep displacing the oil out until the
oil is trapped by water and can no longer be produced. The oil saturation at this
stage is referred to as the residual oil saturation (Sor).

Forced displacement behavior is different depending on the wettability of


the rock. In addition, capillary pressure is also a function of saturation history
(hysteresis), as shown in Figure 8.8.

For a water-wet system, the displacement is simple: the water will flow in
the narrow throats, leaving the oil stranded in the large pores. The residual oil
saturation can be achieved in the spontaneous imbibition phase as the system
prefers to be in contact with water and thus will soak the water in spontaneously.
Based on the capillary pressure curve, all of the water saturation will increase in
the spontaneous imbibition regime and no further increase will be achieved in
the forced injection regime (Figure 8.9).

For an oil-wet system, there will be no oil recovered in the spontaneous


water imbibition regime as the system does not like the water and thus will not
soak it . Then, there will be a significant recovery by forced water injection (Figure
8.9).

For a mixed-wet system, both oil-wet layers and water-wet layers can lead to
medium oil recovery by spontaneous imbibition, followed by significant further
recovery by forced water injection as a greater force is applied to displace the oil
(Figure 8.9). However, mixed-wet systems are very complicated as the ratio of
water-wet to oil-wet fractions in a rock can vary significantly, which can lead to a
totally different behavior from the one displayed.

The shape of the capillary pressure curve can help in identifying the wettability
state of the system as it is a function of wettability. In addition, the concept of
Amott index covered in Chapter 7 can now be revisited as all the parameters
discussed in Equations 7.4 and 7.5 can now be obtained from capillary pressure
curves as shown in Figure 8.10.

158
Swir Sor
Irreducible water Residual
oil
(5) Secondary drainage
(forced)
Pc (+ve)
(1) Primary drainage

(2) Imbibition of water


(spontaneous)

Pc (-ve) (3) Water injection (forced)


(4) Oil invasion (spontaneous)

0 Sw 1

Figure 8.8: Schematic showing the different cycles of capillary pressure and fluid saturation
notations.

+
Water-wet
Mixed-wet
Oil-wet

Pc 0

-
0 0.2 0.4 0.6 0.8 1
Sw

Figure 8.9: Schematic showing the effect of wettability on water re-saturation capillary pressure
curves.

159
SW (A) So (E)

Pc (+ve)

SW (B) So (D)

Pc (-ve)

SW (C) So (C)

0 SW 1
Figure 8.10: Schematic showing all the parameters listed in the Amott index equation and their
location in the capillary pressure curves.

Example 8.1
The figure below shows capillary pressure data during primary drainage,
water re-saturation, and secondary drainage for a reservoir rock under
reservoir conditions.

Calculate the Amott wettability index and identify whether the rock is water-,
intermediate-, or oil-wet?
12
Primary drainage
10
Water re-saturation

8 Secondary drainage

6
4
Pc [kPa] 2
0
-2
-4
-6
-8
0 0.2 0.4 0.6 0.8 1
SW

160
Solution

Based on the figure:

Sw(A) = 0.2 Sw(B) = 0.42 Sw(C) = 0.86 So(C) = 0.14 So(D) = 0.4 So(E) = 0.77

Based on the previous observation, we can calculate the following using


Equations 7.4–7.6:
Sw (B) − Sw (A) 0.42 − 0.2
Iw = = = 0.333
Sw (C) − Sw (A) 0.86 − 0.2
So (D) − So (C) 0.4 − 0.14
Io = = = 0.413
So (E) − So (C) 0.77 − 0.14

IA = Iw − Io = 0.333 − 0.413 = -0.08


This indicates that the rock is intermediate-wet according to Table 7.4.

8.3 Laboratory Measurements of Capillary Pressure

Capillary pressure laboratory measurements are part of SCAL and can be


measured using the Porous Plate (PP) technique, Mercury Injection Capillary
Pressure (MICP), and centrifuge.

8.3.1 Porous Plate Technique (PP)

We use a porous plate (which is a low permeability water-wet ceramic disc, this
disc will retain the oil in the core and will let the water pass through the outlet)
to distinguish between the two phases (Figure 8.11). The inlet pressure will be
the oil pressure and the outlet pressure will be the water pressure. Capillary
pressure is the difference between the non-wetting phase and the wetting phase
pressures. In this system, we have a water-wet core sample; thus, the capillary
pressure will be the oil pressure minus the water pressure. We can start with a low
capillary pressure and wait until we reach equilibrium, which is achieved when
no further water is produced, and then we can calculate the water saturation by
weighing or using volumetric balance. We increase the capillary pressure in steps
until we reach the irreducible water saturation. Measurements using the porous
plate technique are very time-consuming, depending on the capillary pressure
selected and the permeability of the core. It may take few hours to several weeks
for one data point. However, the porous plate is the most accurate method
among other capillary pressure measurement techniques, which can also use
reservoir fluids.

161
0 00
30

Po Pw
Gas

00
30
0 00
30

Oil Water
Core
Back Pressure
Regulator

Porous Plate

Figure 8.11: Schematic showing the porous plate experimental set-up to measure capillary
pressure. The back pressure regulator is placed at the outlet to control the outlet pressure.

8.3.2 Mercury Injection Capillary Pressure (MICP)

In this technique, the core sample is placed inside a pressure chamber and
mercury is injected through a pump to the core sample (Figure 8.12). Mercury
acts as the non-wetting phase while air acts as the wetting phase. We inject
mercury to the core and calculate the volume of mercury injected as well as the
pressure. The volume of mercury injected can be converted to a non-wetting
phase saturation. The wetting phase saturation will be one minus the non-wetting
phase saturation. The number of data points will depend on how frequently
the measurements are taken. MICP is the fastest method to measure capillary
pressure experimentally; however, it does not use reservoir fluids and will
require an additional step of conversion that will be addressed in the following
section. Moreover, introducing mercury to a core sample damages the core,
making it unusable for further analysis. Finally, MICP is not fully representative
of the reservoir condition, as mercury can saturate 100% of the core sample
without obtaining a representative irreducible saturation as in reservoirs. MICP
data are usually used to characterize the rock samples by obtaining the throat
radius distribution, which will be covered in the following section.

162
0 00
30

Pressure
Oil in

Valve Closed

Pressure Chamber

Mercury
Core
Sample

(a) (b)

0 00 0 00
30 30

Valve Opened

Increase in
Further Increase
Oil Pressure
in Oil Pressure

(c) (d)

Figure 8.12: Schematic showing the experimental set-up and procedure of the mercury injection
capillary pressure (MICP).

8.3.3 Centrifuge

In this method, we can measure the drainage and waterflood capillary pressure
curves as shown in Figure 8.13. For drainage, we place a core sample filled
with the wetting phase in a cell surrounded by the non-wetting phase (Figure
8.13a). We centrifuge the cell at several speeds starting from low to high. At each
speed, we measure the amount of fluid displaced during the process until we
see no further production at high speeds. At each point, we convert the rotation
speed to pressure using specific equations. For waterflood capillary pressure
curve, the cell will be reversed compared to drainage as the oil phase is usually
less dense than water (Figure 8.13b). The centrifuge method is considered as
an intermediate method between the porous plate and the MICP. Moreover, in
the centrifuge method, we can use reservoir fluids, and the time required to
generate a capillary pressure curve lies between that for the MICP and porous
plate method. However, the accuracy of the method is not as high as in the
porous plate method.

163
Centrifuge Arm

Teflon Cap

Trunion Ring

Graduated
Glass Tube
Core

Centrifuge
Shield Rubber Stopper
Rubber Cushion

(a) (b)

Figure 8.13: Schematic of the centrifuge laboratory method where (a) is the position to measure
water displacing oil capillary pressure while (b) is the position to measure the oil displacing
water capillary pressure.

8.4 Capillary Pressure Conversion and Throat Radius Distribution

Capillary pressure can be measured using any fluid pairs which makes it easier to
measure in the laboratory. In the case of MICP, a conversion needs to be made to
make the curve more representative of the reservoir condition.

First, the capillary pressure in the laboratory for a certain core sample will be the
following:

2σlab cos θlab


Pc,lab = (8.18)
r
where Pc,lab is the capillary pressure measured in the laboratory condition using
specific fluid pairs, and σlab and θlab are the interfacial tension and the contact
angle, respectively, in the laboratory conditions.

Now, if the same core sample is used to measure the capillary pressure using
reservoir conditions, the capillary pressure in this case will be:

2σres cos θres


Pc,res = (8.19)
r
where Pc,res is the capillary pressure measured using reservoir fluids and
conditions, and σres and θres are the interfacial tension and the contact angle,
respectively, in reservoir conditions.

Analyzing Equations 8.18 and 8.19, the radius is the same in both cases since we
use the same core sample.
164
Therefore, we can equate both equations with respect to r to obtain:

σlab cos θlab σres cos θres


= (8.20)
Pc,lab Pc,res

Now we can rearrange the equation to make Pcres as the subject and thus:

σres cos θres


Pc,res = Pc,lab (8.21)
σlab cos θlab

This equation is commonly used to convert capillary pressure to the desired


condition.

MICP can also be used to characterize the rock as it can measure several
capillary pressure data points in a short span of time. These capillary pressure
points can be converted to throat radii if the interfacial tension and contact
angle are known by rearranging Equation 8.17 and making r as the subject.
The throat radii are listed and a frequency of each throat radius is found and
then a plot of the throat radius as a function of the frequency is made, which
is commonly referred to as the throat radius distribution. For example,
Figure 8.14 shows MICP curves for three rocks, two of them being homogeneous
rock samples as they show stable or uniform change in capillary pressure with
respect to water saturation. The difference between the two is mainly the
permeability, with the higher permeability rock having lower threshold capillary
pressure as explained before. The heterogeneous rock was assumed to be the one
due to the instable change in capillary pressure with respect to water saturation.
This can be confirmed further by plotting the throat radius distribution (Figure
8.15). Homogeneous rock samples tend to have unimodal distribution with one
peak, while heterogeneous rock samples tend to have bimodal distribution with
two peaks, or in some cases trimodal with three peaks. This indicates that the
rock has different groups of radii present in the core sample, which makes the
rock heterogeneous.

100
Low Permeability, Homogeneous Rock

High Permeability, Homogeneous Rock

10 Heterogeneous Rock

1
Pc [MPa]

0.1

0.01

0.001
0 0.2 0.4 0.6 0.8 1
Sw

Figure 8.14: Different drainage capillary pressure curves based on different rock properties from
MICP measurements.

165
2.5
Low Permeability, Homogeneous Rock

High Permeability, Homogeneous Rock


2
Throat Radius Distribution

Heterogeneous Rock

1.5

0.5

0
0.001 0.01 0.1 1 10 100
r [µm]

Figure 8.15: Throat radius distribution against throat radius for the rocks shown in Figure 8.14.

Example 8.2
A well is drilled on a structure, and three rock types have been identified
from coring the well (A, B, C). Capillary pressures were derived from the air/
mercury system and converted to height above the free water level.
500

450
Height above free water level [ft]

400

350
A B C
300

250

200

150

100

50

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Sw

Answer the following questions:

a) Which of the three rock types is the “best” reservoir quality rock? Why?

b) Which of the three rock types is the “worst” reservoir quality rock? Why?

c) What is the irreducible water saturation for each of the three rocks?

166
d) Which rock type has the shortest transition zone?

e) Which rock type has the longest transition zone?

f) What is the FWL if the depth of the OWC for rock A was found at 5000 ft
from wireline logging?

g) What is the depth of the OWC for rocks B and C?

h) Assume that a second well is drilled nearby the structure and the same
rock types are identified as rocks (A, B, and C). From wireline logging, it was
found that the well intersects rock type B at a depth of 4000 ft with a water
saturation of 35%.

I. What is the depth of the FWL on this structure?

II. At what depth can we expect to find OWC in each rock type on
this structure?

Solution

a) Rock A has the best reservoir quality since it has the highest permeability.

b) Rock C has the worst reservoir quality since it has the lowest permeability.

c) Rock A: 0.1
Rock B: 0.2
Rock C: 0.1

d) Rock A, since there is little change in capillary pressure as a function of


water saturation, indicating homogeneous pore throats.

e) Rock C, since there is a rapid change in capillary pressure as a function of


water saturation, indicating heterogeneous pore throats.

f) 5000+50 = 5050 ft

g) Rock B: 5050 -70 = 4980 ft


Rock C: 5050 – 100= 4950 ft

167
h) I. 4000 +275 = 4275 ft

II. Rock A: 4275 -50 = 4225 ft


Rock B: 4275 -70 = 4205 ft
Rock C: 4275 -100 = 4175 ft

Example 8.3
Capillary pressure measurement on a reservoir rock in the laboratory gave
the following results:

Sw [-] Pc [bar] Lab conditions Reservoir conditions

0.16 2.4 σ o/w [N/m] 0.05 σ o/w [N/m] 0.03


0.16 1.0
θ o/w 0° θ o/w 25°
0.20 0.75
ρo [kg/m3] 720
0.35 0.5
ρw [kg/m3] 1070
0.85 0.25
0.96 0.1
1 0.05

Based on a wireline resistivity log data from the same reservoir, the OWC
was found at a depth of 1530 m below sea level.

a) Does the data represent drainage or imbibition capillary pressure? Why?

b) What is the irreducible water saturation? Why?

c) Convert the Pc data from laboratory conditions to reservoir conditions.

d) Convert the Pc from part c to height above the free water level.

e) Based on your data, what would be the expected depth of the FWL?

f) What is the thickness of the transition zone?

g) Where would you perforate (create communication between the reservoir


and the well) in the reservoir in order to produce clean oil (oil only)? Why?

168
Solution

a) The data represents a drainage process, as there is a decrease in the


wetting phase, and the entry pressure starts at Sw =1.

b) The irreducible water saturation is 0.16. It can be noted from the table
that as the capillary pressure increases beyond 1 bar, the water saturation
value is not affected.

c) First, we have to convert the capillary pressure to Pa, and then use
Equation 8.21:
σres cos θres
Pc,res = Pc,lab
σlab cos θlab

Sw Pc [bar] Pc [ Pa] Pc,res [Pa]

0.16 2.4 240000 130508


0.16 1.0 100000 54378
0.20 0.75 75000 40784
0.35 0.5 50000 27189
0.85 0.25 25000 13595
0.96 0.1 10000 5438
1 0.05 5000 2719

d) To convert Pc to height, we use Equation 8.16:


Pc,res
h=
(ρw − ρo )g
Sw Pc,res [Pa] h [m]

0.16 130500 38.01


0.16 54400 15.84
0.20 40800 11.88
0.35 27200 7.92
0.85 13600 3.96
0.96 5440 1.58
1 2720 0.79

e) OWC is located at 1530 m, where the water saturation is maximum.


Thus, OWC corresponds to the lowest capillary pressure with Sw = 1,
2720 Pa with a height above the free water level of 0.79. This means that
FWL = OWC + h = 1530 + 0.79 = 1530.79 m.

169
f) The thickness of the transition is: thickness = 15.84 – 0.79 = 15.05 m.

g) We should perforate above the transition zone. At this depth, there is


no free water in the reservoir. The only water existing is the water trapped
inside the pores and this water is irreducible; thus, it does not affect the oil
production. In the transition zone, there will be a mixture of oil and water
produced.

8.5 Leverett J-Function

Capillary pressure can be expressed in a dimensionless form, which accounts


for different interfacial tensions and rock properties known as the Leverett
J-function. This expression is used to convert capillary pressure data that may
be performed with different rock or fluid properties than the field data. For
example, if the capillary pressure was performed in the laboratory on a core
sample with a specific porosity and permeability and the average values for the
same properties are different in the reservoir scale, then the J-function can be
used to account for that.

The Leverett J-function, in SI units, can be written as follows:



Pc k
J(Sw ) = (8.22)
σcos θ φ

where J(Sw) is the dimensionless J-function, Pc is the capillary pressure [Pa],


σ is the interfacial tension [N/m], θ is the contact angle [°], ф is the porosity
[dimensionless], and k is the permeability [m2].

In addition, the J-function in oilfield units is shown below with the conversion
constant:

Pc k
J(Sw ) = 0.21645 (8.23)
σcos θ φ

where J(Sw) is the dimensionless J-function, Pc is the capillary pressure [Psia], σ


is the interfacial tension [dyne/cm], θ is the contact angle [°], ф is the porosity
[dimensionless], and k is the permeability [mD].

Figure 8.16 shows an example of several capillary pressure curves measured


and how the J-function is used to find the average capillary pressure. The process
is very easy and simple: first, find the J(Sw) for the rock used in the laboratory.
Then, use the following rearranged J-function equation to find the new capillary

170
pressure while inputting the J(Sw) found in the first step and the new rock and/or
fluid properties:

φ
Pc = σcos θ J(Sw ) (8.24)
k
This process should be sufficient to convert the capillary pressure to the desired
condition.

J (Sw) 2

0
0 20 40 60 80 100
Sw [%]

Figure 8.16: Schematic showing how the J-function is used to find the desired capillary pressure.
The black line represents the average capillary pressure of the system.

Example 8.4
A core was extracted from a producing oil reservoir, with a porosity and
permeability of 12% and 50 mD, respectively. The capillary pressure-
saturation data were obtained experimentally and found to be related
through the following equation:

Pc = 0.5Sw-0.78

The interfacial tension is measured at 40 dyne/cm with a contact angle of 0°,


and the irreducible water saturation is 0.2.

After measuring the porosity and permeability on several samples, the


average porosity and permeability of the reservoir was found to be at 16%
and 85 mD, respectively.

Use the previous information to generate a representative capillary pressure


data for the reservoir.

171
Solution

First, we have to calculate the capillary pressure at different saturations and


the value of J-function.

Equation 8.23 is used to calculate the J-function J(Sw):



Pc k
J(Sw ) = 0.21645
σcos θ φ

The results are tabulated below:

Sw Pc [psia] J(Sw)

1 0.5 0.055
0.8 0.6 0.066
0.6 0.74 0.082
0.4 1.02 0.11
0.2 1.75 0.19

Then, using the new porosity and permeability value, we find a new
relationship to relate capillary pressure and J-function by rearranging
Equation 8.23.

σcos θ φ
Pc = J(Sw )
0.21645 k
Finally, we reconstruct the capillary pressure-saturation table:

Sw J(Sw) Pc [psia]

1 0.055 0.441
0.8 0.066 0.529
0.6 0.082 0.657
0.4 0.11 0.882
0.2 0.19 1.52

8.6 Water Saturation Distribution in a Layered System

Figure 8.17 shows a reservoir composed of three different rock types. Each rock
type has its own capillary pressure; thus, the water saturation in these layers will
follow in the path of capillary pressure curves, as shown in the figure, as long as
there is no discontinuation between the layers due to impermeable layers. This
concept is very important as now the OWC will be in the layer closest to the FWL.
In addition, the oil saturations will vary in each layer depending on the capillary
pressure, and assuming one capillary pressure for the entire reservoir might lead
to errors in estimating oil saturations.
172
Lithology

Sw (h)

B Pc A C B

0 Sw 1

Figure 8.17: Schematic showing a reservoir with three layers of different rocks and how fluid
saturation will vary in each layer depending on the capillary pressure curve.

8.7 Hydrostatic Pressure and Repeat Formation Tester (RFT)

As discussed earlier, hydrostatic pressure is the fluid pressure under static


conditions, repeat formation tester (RFT) is a tool that measures hydrostatic
pressure as a function of depth on the borehole wall. Reservoir fluid samples
can also be taken from the tool as well for further analysis on the surface.

From the RFT tool, the free levels (oil and water) can be identified, as well as the
density of the fluids present in the reservoir. It can also give an indication of the
reservoir pressure status.

For the free water identifications, let us assume that we have a reservoir as
shown in Figure 8.18 which has gas, oil, and water. The pressure response is also
shown in the same figure, which indicates that the slope changes as the density
of the fluid changes (P = ρgh). At the intersection point between the gas and oil or
oil and water, the pressures of each phase are equal to each other, that is, at the
intersection of gas and oil (Pg = Po) and the same for oil and water (Po = Pw). Thus,
the capillary pressure at these points is zero, which indicates the free oil level
(FOL) and free water level (FWL) for each section.

By using hydrostatic pressure for two points, one in each phase, the free levels
can be identified.
173
Free Oil Level

Let us assume that we have one measured pressure point from the RFT in the
gas reservoir Pg at a depth of Zg from the surface level and similarly one point for
the oil Po at a depth Zo as shown in Figure 8.18. In this case, we can say that the
pressure of the FOL from the gas side is equal to:

PZF OL = Pg + ρg g(ZF OL − Zg ) (8.25)

Similarly, the pressure of the FOL from the oil side is equal to:

PZF OL = Po − ρo g(Zo − ZF OL ) (8.26)

Note that the pressure increases as we go downward and decreases as we go


upward.

Now, equating both equations and solving for ZFOL, which is the depth of the free
oil level, yields:

Po − Pg + ρg gZg − ρo gZo
ZF OL = (8.27)
(ρg − ρo )g

Free Water Level

Finding the depth of the FWL is similar to the ZFOL for two data points in each
region (oil and water); we first start by defining pressure from the oil side:

PZF W L = Po + ρo g(ZF W L − Zo ) (8.28)

Similarly, the pressure of the ZFWL from the water side is equal to:

PZF W L = Pw − ρw g(Zw − ZF W L ) (8.29)

Now, equating both equations and solving for ZFWL, which is the depth of the free
water level, yields:

Pw − Po + ρo gZo − ρw gZw
ZF W L = (8.30)
(ρo − ρw )g

In addition to identifying the depth of the free levels, we can identify the density
of the fluids present in the reservoir. The gradients in the RFT data indicate a
change in fluid density and thus from the hydrostatic pressure equation, we can
estimate the density. However, it is better to find the slope of several data points,
in order to find the average density across the zones.

174
RFT data can also indicate the status of the reservoir pressure. For example, we
can compare the water pressure value from the RFT to the theoretical value (Pw
= PA + ρwgZw) of the same depth where the measurement is taken. We chose the
water pressure because water is the only continuous phase in the entire reservoir
due to the presence of the irreducible water saturation in all the zones. The
discrepancy between the theoretical and measured values from RFT can lead to
three conclusions. First, the reservoir is normal-pressured when the theoretical
and measured values are the same. The reservoir is under-pressured when the
RFT data is lower than that of the theoretical one. This indicates that the flow
will not be as rapid as expected. The reservoir is over-pressured when the RFT
data is higher than the theoretical value. This indicates a higher flow rate than
expected, which can lead to blowout if no measures are taken.

Pressure

( Pg , Zg ) Gas
Free oil level (FOL)

( Po , Zo ) Depth
Oil
Free water level (FWL)
( Pw , Zw )
Water

Zg
( Pg , Zg )
= ZFOL - Zg
ZFOL
( Po , Zo ) Zo = Zo - ZFOL

( Pw , Zw )

( Pg , Zg )

Zo
( Po , Zo )

= ZFwL - Zo
ZFwL
( Pw , Zw ) Zw = Zw - ZFWL

Figure 8.18: Schematic showing a reservoir with gas, oil, and water zones and the excepted RFT
response for these zones. In addition, a schematic is made to illustrate how to find the depths of
the free levels (ZFOL and ZFWL).

175
Example 8.5
The following pressure measurements are made for an gas/oil reservoir:

Depth [m] Pressure [MPa] Fluid and density [kg/m3]

2,690 15.12 Gas, 410


2,720 15.33 Oil, 820
2,850 16.41 Water, 1,040

a) Find the depths of the FWL and FOL.

b) Is the reservoir over-pressured, normal-pressured, or under-pressured?

Solution
a) We will use Equations 8.27 and 8.30 to find the depths of the free levels:
Po − Pg + ρg gZg − ρo gZo Pw − Po + ρo gZo − ρw gZw
ZF OL = ZF W L =
(ρg − ρo )g (ρo − ρw )g

which yields:

(15.33 − 15.12) × 106 + (410 × 2690 − 820 × 2720)9.81


ZF OL = = 2697.7 m
(410 − 820)9.81

(16.41 − 15.33) × 106 + (820 × 2720 − 1040 × 2850)9.81


ZF W L = = 2833.6 m
(820 − 1040)9.81

b) Under-pressured, because Pw = PA + ρwgh = 101,000 + 1040 × 9.81 × 2850 =


29.4 MPa, which is higher than the actual water reservoir pressure of 16.41
MPa.

8.8 Applications of Capillary Pressure

Drainage and water re-saturation capillary pressure curves have different


applications. Drainage capillary curves are used to find the fluid saturation above
the free water level. These curves can be used to find the actual depth of the FOL/
FWL from the RFT data, as explained before, or to find the gas/oil contact and oil/
water contact GOC/OWC from wireline logging, as briefly discussed in Chapter
6. Once a depth is identified, the capillary pressure curve can be calibrated to
find fluid saturations at specific depths. In addition, it can identify the depth of

176
the transition zone and the clean oil zone. Water re-saturation capillary curves,
on the other hand, tell us about the flow of oil when water is introduced in the
system. They also can help in identifying the wettability of the system. However,
they are not frequently used in the petroleum industry because (1) they are
difficult to be measured in the laboratory and that is because the spontaneous
imbibition process is rapid and difficult to be measured accurately, and (2) relative
permeability (which will be discussed in Chapter 9) is easier to be measured and
gives us more information about multi-phase flow.

8.9 Summary

Capillary pressure is the pressure difference between the non-wetting phase


and the wetting phase across the interface. Capillary pressure can be divided
into drainage and water re-saturation. Drainage capillary pressure curves are
commonly used to estimate the fluid saturation above the free water level
while water re-saturation curves are used to characterize the efficiency of water
injection to displace oil. Capillary pressure can be measured in the laboratory
using the porous plate, mercury injection capillary pressure, and centrifuge
method. The porous plate is the superior method in terms of accuracy; however,
it is the most time-consuming method. Capillary pressure can be converted to
account for different fluid and rock properties using the J-function. Table 8.2
summarizes the concept of capillary pressure.

Table 8.2 Summary of fluid saturation and its importance to the petroleum industry.

Parameter Symbol Definition Importance

Capillary Pc The pressure We use drainage


Pressure difference capillary pressure
between the non- to determine fluid
wetting phase distribution in
and the wetting reservoir (initial
phase across the condition) above the
interface. free water level.

177
End of Chapter Questions

Question 8.1

An air-brine capillary pressure experiment on a core sample gives the following


results:
Capillary 0 68.95 93.08 96.53 199.95 330.95 689.48
pressure [kPa]

Water 100 100 89 79.6 42.5 34.1 28


saturation [%]

a) Plot water saturation versus height using the data given above.

b) Find the water saturation of a sample that was extracted


50 ft above the OWC.

Given:
σ cos θ a/w [N/m] 0.072

σ cos θ o/w [N/m] 0.03

ρw [kg/m3] 1028

ρo [kg/m3] 689

Question 8.2

The following formation tester pressure measurements were made in an


exploration well. Estimate the following:

a) FOL and FWL


b) The density and the nature of the fluids present in the reservoir

Depth [m] Pressure [MPa]

2474.98 20.08
2499.97 20.13
2524.96 20.18
2549.96 20.34
2574.95 20.53
2599.94 20.72
2624.94 20.95
2649.93 21.20
2674.92 21.47

178
Question 8.3

A set of capillary pressure data for an air/water system (laboratory conditions) is


given in the table below:

Pc [Pa] Sw Lab conditions Reservoir conditions

3,447 1 σ a/w [N/m] 0.072 σ o/w [N/m] 0.03

6,205 0.8 θ a/w 0° θ o/w 30°

16,892 0.6

48,263 0.4

999,740 0.3

a) What is the flow process for this capillary pressure curve:


drainage or imbibition? Explain?

b) Convert the capillary pressure to oil/water (reservoir conditions)


system and plot it using a semi-log scale.

c) On the same figure, label the following: FWL, OWC, Swir.

Question 8.4

The capillary pressure data is given below:


Pressure [Pa] Sw

200,000 0.30
100,000 0.35
10,000 0.45
0 0.58
-10,000 0.6
-100,000 0.6
-200,000 0.6

a) What is the flow process for this capillary pressure curve:


drainage or imbibition? Explain.

b) Identify the wettability of the core:


is it water-wet, mixed-wet, or oil-wet? Explain.

c) Find the irreducible water saturation and the residual oil saturation.

179
Question 8.5

Measure the following primary drainage capillary pressure in the laboratory


using mercury. The core has a permeability of 5.92 × 10-13 m2, a porosity of 0.20,
its interfacial tension is 480 mN/m, and contact angle is 40o.

Pressure [Pa] Saturation

0 1
50,000 1
74,000 0.6
150,000 0.4
250,000 0.3
300,000 0.3

In an oil reservoir, some parameters were measured. The average permeability


is 1.97 × 10-13 m2, the porosity is 0.15, the interfacial tension is 25 mN/m, and the
contact angle is 0o. Plot a graph of water saturation against height above the free
water level in the reservoir.

Question 8.6

The following are RFT data:

Depth [m] Pressure [MPa] Fluid and density [kg/m3]

2,250 17.51 Gas, 305

2,285 17.63 Oil, 650

2,327 18.01 Water, 1,040

Depths are measured from the surface.

a) Find the depths of the free oil and free water levels.

b) Is the reservoir under-pressured, over-pressured,


or normal-pressured?

180
181
182
Chapter 9

Relative
Permeability
When dealing with more than one fluid in a porous medium, the concept of
single-phase permeability is no longer valid to characterize the flow in that
system and thus an extension of Darcy’s law is required. Relative permeability
is defined as the relative ease for one fluid to flow in the presence of another
fluid. A hydrocarbon reservoir, as discussed before, can have oil and water, gas
and water, and gas, oil, and water present at its pore spaces. Therefore, relative
permeability is considered to be a transport property that helps in estimating oil
production when another fluid is injected. Moreover, relative permeability is a
function of pore geometry, saturation history, wettability, and fluid saturation ,
which is mainly used to study the effectiveness of waterflooding as a displacement
process in reservoirs, for example. In addition, laboratory measurements of
relative permeability are part of SCAL.

When discussing relative permeability, three parameters need to be discussed:

1) Relative permeability (kr), which is a measure of the relative ease of


one fluid to flow in the presence of another immiscible fluid.

2) Absolute permeability (k), which is a measure of the ease of one fluid


to flow in a porous medium.

3) Effective permeability, which is the fluid conductance capacity of a


porous medium to a particular fluid when more than one fluid is present
in the porous medium. kw, ko, kg are the effective permeability of water,
oil, and gas, respectively.

In terms of equations:

The relative permeability of water (krw) is defined as the effective permeability of


water (kw) divided by the absolute permeability (k):

kw
krw = (9.1)
k
Similarly, the relative permeability of oil (kro) is defined as the effective permeability
of oil (ko) divided by the absolute permeability:

ko
kro = (9.2)
k

183
To estimate the flow rate when more than one fluid is present in the porous
medium, we need to extend Darcy's law from single-phase flow, as discussed in
Chapter 4, to multiple phases by incorporating Equations 9.1 and 9.2 to replace
the single-phase permeability in Equation 4.8 for linear flow as follows:

kA∆P
q= (9.3)
µL
where q is the flow rate [m3/s], k is the absolute permeability [m2], A is the cross-
sectional area [m2], dP is the difference between the inlet pressure and the outlet
pressure [Pa], μ is the viscosity [Pa.s], and L is the core length [m].

Then, Darcy's law for two-phase flow can be written as:

kw A kkrw A
qw = ∆Pw = ∆Pw (9.4)
µw L µw L

ko A kkro A
qo = ∆Po = ∆Po (9.5)
µo L µo L

where kr is the relative permeability, and subscripts w and o denote water and oil,
respectively. Note that the same concept is applicable for radial flow.

9.1 Relative Permeability Curves

Relative permeability curves can also be divided into two types: drainage and
water re-saturation (which is commonly referred to as imbibition in the literature).
Figure 9.1a is a drainage relative permeability curve where the wetting phase
decreases. In general, drainage relative permeability curves represent the
process of oil migration from the source rock into a water-wet reservoir. The
reservoir or rock is first fully saturated with water, which means that the effective
permeability of water and absolute permeability are the same since there is
no additional fluid in the system, and thus the water relative permeability is
1. As the oil gradually enters the system, the water saturation decreases and
the oil saturation increases; this explains the gradual decrease in the water
relative permeability and the increase in the oil relative permeability. As the oil
saturation increases, the oil flow will increase, while the water saturation which
obstructs the oil flow decreases. The oil relative permeability will increase until
it displaces most of the water out of the system and will reach the irreducible
water saturation where water can hardly be reduced any further. At this point,
only oil will flow and the expected oil relative permeability will be very high (1
or approximately 1), which indicates that the irreducible water saturation will
not contribute significantly to the flow of oil when compared to the single-phase
permeability. However, if the irreducible water blocks some oil flow pathways,
then the maximum oil relative permeability (or end point oil relative permeability)
will be lower than 1.

184
Drainage relative permeability represents the current status when the oil reservoir
is discovered and does not hold as much significance to the petroleum industry
as water re-saturation relative permeability curves. Water re-saturation relative
permeability curves occur after drainage when water is injected in the reservoir
to displace the oil from it a certain stage of production. The water re-saturation
relative permeability curves will start from where they ended in the drainage
relative permeability curves (Figure 9.1b). As water is injected into the system,
the water saturation will gradually increase, starting from the irreducible water
saturation, while the oil saturation and oil relative permeability will decrease. The
water relative permeability will keep increasing until the oil becomes trapped in
the pore spaces (for the water-wet case), known as the residual oil saturation , as
shown in Figure 9.2. The maximum water relative permeability (or the end point
water relative permeability) will be at the residual oil saturation. The shape and
end point water relative permeability is primarily a function of wettability, which
will be discussed next.

1 1

0.8 0.8

0.6 0.6
Oil Water
kr kr
0.4 0.4
Oil Water
0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Sw Sw

(a) (b)
Figure 9.1: Schematic showing (a) drainage relative permeability and (b) water re-saturation
relative permeability for the oil/water system.

185
1

kr kro

krw

0
Swir Sor

0 Sw 1

Figure 9.2: Schematic showing water re-saturation (waterflooding) relative permeability with the
important labeling.

Similar to capillary pressure, relative permeability is also a function of wettability.


Therefore, the wettability of a rock can be indirectly measured by examining
water injection relative permeability curves. Water-wet curves (Figure 9.3a) are
characterized by low water relative permeability as a function of water saturation
as well as low end points (< 0.2); this is because water-wet surfaces prefer to be
in contact with water and moving water out will be harder, which explains the
low mobility of water. Another feature that characterizes water-wet curves is the
crossover of water and oil relative permeability. The crossover of such systems
usually occurs at a water saturation of > 0.5. Moreover, the crossover indicates
that both oil and water flow at equal ease. The residual oil saturation can also
help in identifying the wettability as water-wet rocks tend to have higher residual
oil saturation when compared to oil-wet rocks, which will be explained below.

Oil-wet water re-saturation curves (Figure 9.3b) act in the opposite manner
when compared to water-wet systems. First, the oil relative permeability will be
low while the water relative permeability will be high as the surface wants to be
in contact with oil rather than water, and thus the water will have higher mobility
when compared to oil. Therefore, the water relative permeability end-point will
be at maximum. The crossover will occur at a water saturation of < 0.5 and the
residual oil saturation tends to be lower than that of the water-wet system, which
will also be explained in more detail in the following section.

Now we will analyze water-wet and oil-wet systems from a pore-scale perspective.
Figure 9.4 shows the process of injecting water into a water-wet system to
displace oil. Since the system is water-wet, the water prefers to be in contact
with the surface and thus fills the narrow pores while the oil fills the large pores,

186
giving the water poor connectivity. For the same reason, oil has a higher relative
permeability than water. The residual oil saturation is high though, since oil can
be trapped in the centers of the larger pore spaces by water cornering it from
all sides.

For an oil-wet system (Figure 9.5), we see the reverse of the water-wet scenario
where oil has a lower relative permeability than water since oil will now fill the
narrow pores. Water, on the other hand, will fill the large pores, giving low flow
conductance to oil. However, the oil remains connected in layers until a low oil
saturation is reached.

There are several factors that affect relative permeability, including wettability,
fluid saturation, pore geometry (similar to single-phase permeability), and
saturation history (hysteresis). The effect of hysteresis on relative permeability,
for example, is shown in Figure 9.6 where the relative permeability curve will
vary depending on the flow process.

1 1

0.8 0.8
Oil Oil Water
0.6 0.6
kr kr
0.4 0.4

0.2 0.2
Water
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Sw Sw
(a) (b)

Figure 9.3: Schematic showing waterflood relative permeabilities for (a) water-wet and (b) oil-wet
cases.

Rock Water Oil


Figure 9.4: Schematic showing a pore-scale scenario of water injection into a water-wet rock
where oil is eventually trapped in the middle of the pores surrounded by water leaving the
residual oil behind.

187
Rock Water Oil

Figure 9.5: Schematic showing a pore-scale scenario of water injection into an oil-wet rock where
oil sticks to the surface in the narrow pore spaces which yields poor conductivity for oil.

1
Water re-saturation
Secondary Drainage
0.8
Primary Drainage

0.6

kr
0.4

0.2

0
0 0.2 0.4 0.6 0.8 1

Sw

Figure 9.6: Schematic showing the effect of hysteresis on relative permeability curves.

9.2 Recovery Factor Estimation

Recovery factor (RF) can be defined as the fraction or percentage of the oil produced
when compared to the oil present in the reservoir. From waterflooding relative
permeability, we can estimate the potential recovery factor by waterflooding, for
instance, but first we need to discuss some concepts which include displacement
efficiency (ED) and volumetric sweep efficiency (EV).

9.2.1 Displacement Efficiency (ED)

Displacement efficiency (ED) is the fraction of oil recovered from a zone swept
by water or another fluid. Figure 9.7 shows a waterflood relative permeability
where water displaces oil. This figure can be divided into two parts to estimate
the potential displacement efficiency of a rock or reservoir. The first part is the
available oil in an oil zone which is referred to as the initial oil saturation, Soi= 1-
Swir (the oil present excluding the irreducible water saturation). The other part is

188
the recoverable oil by water injection which is all the oil in the oil zone, excluding
the trapped oil at the pore spaces (residual oil saturation); this would be 1 - Swir -
Sor and hence the displacement efficiency will be:

1 − Swir − Sor
ED = (9.6)
1 − Swir

kr kro

krw

0
Swir Sor

0 Sw (1) 1

Available oil
(2)
Recoverable oil

Figure 9.7: Schematic showing the two parts of displacement efficiency, the available oil (1) and
recoverable oil (2). The ratio of the recoverable oil (2) to the available oil (1) is the displacement
efficiency.

9.2.2 Volumetric Sweep Efficiency (EV)

Injecting water into a reservoir is an efficient method to sweep oil out of the
pore spaces; however, water will not be able to access the entire reservoir
due to the reservoir’s heterogeneity and fluid density. Moreover, water will
take the easiest path from the injection to the production well as shown in
Figure 9.8. The easiest path to flow is usually through the highest permeability
zone as it requires the lowest pressure drop. Therefore, the amount of area
accessed by water with respect to the total horizontal reservoir area is known as
the areal sweep efficiency (EA), while the amount accessed by water, for instance,
189
in the vertical column with respect to the total oil column is known as the vertical
sweep efficiency (Ev). For Ev, the sweep efficiency can vary depending on the
permeability of the reservoir and also on the type of fluid injected. For example,
injecting water into an oil reservoir will cause the water to sink downward as
water is denser than oil, while gas will flow upward in an oil reservoir because
the gas is less dense and thus accesses a smaller portion of the reservoir. The
volumetric sweep efficiency (EV) is the product of the areal sweep efficiency and
the vertical sweep efficiency, as shown in the equation below:

EV = EA .Ev (9.7)

Producer

EA

Injector Soi

Sor

Ev

Figure 9.8: Schematic showing the concept of areal and vertical sweep efficiency in an oil reservoir
where water is used to sweep the oil to the producer well.

9.2.3 Recovery Factor (RF)

The recovery factor estimated from the injection of a particular flood can be
defined as:

RF = ED .EV (9.8)

Figure 9.9 shows the recovery factor as a function of pore volume of water
injected into water-wet and oil-wet systems for arbitrary cases. First, when water
is injected into a reservoir, an equal volume of oil will be produced as the volume
is conserved and that can occur before the water breakthrough to the production
well. Once the water reaches the production well, the oil production will decrease
as both water and oil are produced. Since the water mobility is low for water-wet
systems and high for oil-wet systems, the breakthrough of water occurs faster
in oil-wet systems as water will reach faster to the production well in oil-wet
systems. For water-wet systems, there will be little to no oil production after
water breakthrough as water surrounds the oil, rendering it immobile. However,
for the oil-wet system, oil will form a layer on the small pores as water is injected
190
even after water breakthrough; there will be continued, slow production after
water breakthrough even after 5000 pore volumes of water injected in some
scenarios. Injecting this amount in a reservoir is unrealistic and not economically
feasible, and thus a decision will be made to move to another more efficient
displacement process after injecting few pore volumes of water.

100
Recovery Factor [%]

Water-Wet
Oil-Wet

0
0 Pore Volume Injected 60

Figure 9.9: Schematic showing the recovery factor as a function of pore volume of water injected
into water-wet and oil-wet arbitrary cases.

Example 9.1
Based on the figure below, answer the following questions:

0.8 Oil

0.6
kr
0.4
Water

0.2

0
0 0.2 0.4 0.6 0.8 1
Sw

a) What is the wettability of the rock?

191
b) What is the flow process of the relative permeability curve?

c) What is the value of the irreducible water saturation?

d) What is the value of the residual oil saturation?

e) What is the water relative permeability at the residual oil


saturation?

f) From the same figure, what is the displacement efficiency?

g) What would be the expected recovery factor of a field with the


same relative permeability, as shown in the figure below, if the
volumetric sweep efficiency is 70%?

Solution
a) Water-wet

b) Imbibition

c) 0.15

d) 1 - 0.7= 0.3

e) 0.3

f) Using Equation 9.6, ED= 0.64

g) RF = ED × EV = 0.64 × 0.7 = 0.45 or 45%

9.3 Laboratory Measurement of Relative Permeability

Relative permeability measurement is part of SCAL measurements and can


be generally measured using the steady-state and unsteady-state laboratory
methods.

9.3.1 Steady State (SS)

In general, for the steady-state approach, both fluids are injected simultaneously;
the flow rate and pressure drop are measured to find the relative permeability
(Figure 9.10). Then, we can calculate relative and effective permeability. After
that, the injection rates are changed to obtain a new set of relative permeability

192
points. The time required between each set of injection rates is between a few
hours to days depending on the absolute permeability in order to reach a steady-
state regime and thus be able to obtain a new set of relative permeability points.

Briefly, the first step would be to saturate the core 100% with water to measure
its absolute permeability. Then, oil is injected to drain the water out of the rock
to reach the irreducible water saturation. This condition represents the initial
reservoir condition and where the end point of oil relative permeability (kro,max)
can be measured. After that, both oil and water are injected simultaneously
while water is injected at a low flow rate and oil is injected at a higher rate. The
system needs to reach a steady-state regime before any reading can be taken.
The water saturation is usually measured using volumetric measurement,
weighing, electrical resistivity, or X-ray imaging to monitor the saturation inside
the core. After the system reaches the steady state and the data are recorded,
the water flow rate is slightly increased while the oil flow rate is decreased. The
incremental change in flow rate depends on the time available to perform the
experiment. More data points will require lower incremental changes in flow
rate. The incremental increase in water flow rate will continue until only water is
flowing in the system and thus the residual oil saturation is achieved.

Steady-state relative permeability is considered more accurate than the


unsteady-state method which is covered in the following section; however, the
steady-state method is more time-consuming.

193
Oil
Water 100% Sw
Water

ΔP

Oil

Swir kro,max

Oil
A
Water Swir
A

Oil
B
Water
Swir
B

Oil
C

Water
Swir
C

Sor
D

Water
Swir krw,max

Figure 9.10: Schematic of a steady-state waterflood relative permeability experiment.

194
Several models can be used to express relative permeability curves that are
mainly used to smooth experimental data or represent relative permeability in
geological models. The most commonly used models are the Corey models:

 n
Sw − Swir
krw = krw,max (9.9)
1 − Sor − Swir
 m
1 − Sw − Sor
kro = kro,max (9.10)
1 − Sor − Swir

where all the parameters in these equations are given in Table 9.1.

Table 9.1: List of all the parameters associated with the Corey curves.

Parameter Definition

krw Water relative permeability

kro Oil relative permeability

krw, max Maximum water relative permeability

kro, max Maximum oil relative permeability

Sw Water saturation (you need to input the water saturation


values between Swir – 1- Sor of your system)

Swir Irreducible water saturation

Sor Residual oil saturation

n Corey's water exponent

m Corey's oil exponent

Example 9.2
Given a rock with 80 mD absolute permeability, 13.2% porosity, 6 in length,
and 1.5 in diameter. The dry weight of the rock is 126 g. During a two-
phase steady state-relative permeability experiment, water was injected at
a rate of 0.23 cm3/min while oil was injected simultaneously at a rate of

195
0.52 cm3/min. The outlet of the rock was open to atmospheric pressure
while the inlet pressure was measured at 7 psig for both water and oil. The
water viscosity is 1 cP and oil viscosity is 0.9 cP. The rock weighs 145.7 g
at that stage of the experiment. The water density is 1 g/cm3 while the oil
density is 0.72 g/cm3.

Based on the information given, calculate:

a) Water and oil effective permeability values.

b) Water and oil relative permeability values.

Solution
a) First, the dimensions of the sample need to be converted to cm so that
the units are consistent. Since 1 in = 2.54 cm:

Diameter = 1.5 in = 1.5 × 2.54 = 3.81 cm


Length = 6 in = 6 × 2.54 = 15.24 cm

Then, the cross-sectional area of this core is:

d2 3.812
A = π 4 = π 4 = 11.4 cm
2

After that, we have to make sure that the equation is dimensionally


consistent before we proceed; thus, we convert the rates from cc/min to
cc/s, yielding qw = 3.83×10-3 cc/s and qo = 8.67×10-3 cc/s. Also, the pressure
P1 at the inlet is converted from 7 psig to 1.476 atm. Using Darcy’s law in
Equations 9.4 and 9.5 and rearranging the parameters, we can calculate
water effective permeability and oil effective permeability, respectively,
given by:
qw µw L 3.83 × 10−3 × 1 × 15.24
kw = = = 0.0108 D or 10.8 mD
A(P1 − P2 ) 11.4(1.476 − 1)

qo µo L 8.67 × 10−3 × 0.9 × 15.24


ko = = = 0.0219 D or 21.9 mD
A(P1 − P2 ) 11.4(1.476 − 1)

196
b) We use Equations 9.1 and 9.2:

kw 10.8
krw = = = 0.135
k 80

ko 21.9
kro = = = 0.273
k 80

9.3.2 Unsteady State (USS)

In an unsteady-state measurement (also referred to as displacement method),


as shown in Figure 9.11, the rock or the core is saturated with one fluid (water
or oil), and then the other fluid is injected to displace the first fluid. During the
unsteady-state displacement, the time is recorded for flow rates of both fluids.
Applying this method is cheap and relatively quick; however, this method requires
analytical interpretation of the results. That is because relative permeability is
not measured directly, which makes the results less reliable than the steady-
state method.

197
Oil
100% Sw
Water

Oil Oil

Swir kro,max

A
Water Oil

Swir

Water B Oil

Swir
B = Water Breakthrough
B

C
Water B Oil + Water

Swir

Sor
D
Water Water

Swir krw,max

Figure 9.11: Schematic of an unsteady-state waterflood coreflood relative permeability


experiment.

9.4 Three-Phase Relative Permeability

Three phases (water, oil, and gas) can flow simultaneously sometimes and in
order to characterize the flow in these cases, three-phase relative permeability
is introduced. The usage of three-phase relative permeability is minimal in
the petroleum industry when compared to two-phase relative permeability. In

198
addition, laboratory measurement of three-phase relative permeability is very
difficult to perform. Therefore, Stone came up with two models Stone I and
Stone II to estimate three-phase relative permeability from two-phase relative
permeability (oil/water and oil/gas); the details of the two models are not
discussed in this book and can be found elsewhere.

Three-phase relative permeability is presented in ternary diagrams (Figure


9.12) where each phase is placed in an apex with gradual decrease in saturation
away from the respective apex. The water, oil, and gas three-phase relative
permeabilities are shown in Figures 9.13–9.15, respectively.

100% Gas

80%
20% 20%

60% 40%
40%

40%
60% 60%

20%
80% 80%

100% Water 100% Oil

Figure 9.12: A ternary diagram showing water, oil, and gas saturations.

100% Gas

krw 10%

20%

40%

60%

80%

100% Water 100% Oil

Figure 9.13: Schematic showing a water relative permeability displayed in a ternary diagram of
a three-phase system.

199
100% Gas

1%

5%
%
10 kro

20%
30%
40%
50%
60%

100% Water 100% Oil

Figure 9.14: Schematic showing an oil relative permeability displayed in a ternary diagram of a
three-phase system.

krg
100% Gas

50%
40%

30%

20%

10%

5%

1%

100% Water 100% Oil

Figure 9.15: Schematic showing a gas relative permeability displayed in a ternary diagram of a
three-phase system.

Example 9.3
Answer the following questions:

a) Indicate on the figure the point where gas saturation is 20%, oil
saturation is 50%, and water saturation is 30%, label it as A.

200
b) Indicate on the figure the point where oil saturation is 80% and
water saturation is 20%, label it as B.

c) What is the value, in fraction, of water relative permeability at 20%


gas saturation, 40% oil saturation, and 40% water saturation?

100% Gas

krw 10%

20%

40%

60%

80%

100% Water 100% Oil

Solution
100% Gas

krw 10%

20%

40% A

60%

80%
B

100% Water 100% Oil

a) See the plot

b) See the plot

c) krw = 0.03

201
9.5 Summary

When dealing with a system that has two fluids flowing in it, there are two
important parameters to characterize the flow in the system, which are effective
and relative permeability. Relative permeability is the ratio of the effective
permeability, which is the permeability of one fluid in the presence of another
fluid, to the absolute permeability. Several factors affect relative permeability,
which include pore geometry, wetting properties, saturation history, and fluid
saturation. Knowing relative permeability is important in understanding recovery
factor, which is the product of the displacement efficiency and the volumetric
sweep efficiency. There are two ways of measuring relative permeability: (a)
steady-state method or (b) unsteady-state (displacement) method. Steady state
is the most common and accurate method to measure relative permeability. The
concept of relative permeability can extend to a system with three phases. Three-
phase relative permeability is very difficult to be performed experimentally and
usually models are used to portray three-phase relative permeability in ternary
diagrams. Table 9.2 summarizes the concept of relative permeability.

Table 9.2: Definition of relative permeability and its importance to the petroleum
industry.

Parameter Symbol Definition Importance

Relative kr It is the ratio We use relative


Permeability of the effective permeability to study
permeability and understand the
of that phase multi-phase flow
to the absolute in reservoirs. For
permeability. instance, we use
relative permeability
to predict the
recovery of oil by
water injection.

202
End of Chapter Questions

Question 9.1

Use the figure below to answer the following questions:

a) What process does the relative permeability curve show?

b) What is the wettability of the core?

c) What is the irreducible water saturation?

0.8

Oil
0.6
kr

0.4

0.2
Water

0
0 0.2 0.4 0.6 0.8 1
Sw

Question 9.2

Use the figure and the table below to answer the following questions:

a) What is the value of oil effective permeability, in mD, in rock R4 at 45%


water saturation?

b) What is the volume of water, in cm3, recovered from R4 at the irreducible


water saturation?

c) If oil was injected at 1 cm3/min in R4 (to reach the irreducible


saturation), the outlet pressure was maintained at atmospheric
conditions, what is the expected pressure, in atm, at the inlet of the
core?

203
Rock R4

Length [in] 2.76

Diameter [in] 1

Dry weight [g] 86.79

Pore volume [cm3] 2.77

Porosity [%] 7.8

Absolute permeability [mD] 4.67

µw [cP] 1

µo [cP] 0.9

0.8

Oil
0.6
kr

0.4

Water
0.2

0
0 0.2 0.4 0.6 0.8 1
Sw

Question 9.3

The data given below are for an oil reservoir that is composed of a single, fairly
homogeneous rock. Based on the below information, find:

a) The displacement efficiency.

b) The recovery factor.


1

0.8

Oil
0.6
kr

0.4

0.2
Water

0
0 0.2 0.4 0.6 0.8 1
Sw

204
Cross-sectional view of the reservoir after water injection

Top view of the reservoir after water injection

1- Sor Soi= 1- Swir

Question 9.4

Steady-state experiments were conducted on two rocks and the fitted Corey
curve parameters are shown below:

krw,max kro,max Swir Sor n m

Rock A 0.1 0.9 0.1 0.3 4.5 2.5

Rock B 0.8 0.3 0.1 0.05 3.0 5

Based on the data above, answer the following questions:

a) Plot the relative permeability curves for both rocks.

b) What is the wettability of each rock? Why?

c) What is the displacement efficiency of each rock?

d) Assuming that the volumetric sweep efficiency is 0.65, what is the


expected recovery factor?

205
Question 9.5

Answer the following questions based on the figure below:

a) Indicate on the figure the point where gas saturation is 80% and water
saturation is 20%.

b) What is the value, in fraction, of gas relative permeability at 40%


gas saturation, 40% oil saturation, and 20% water saturation?

c) What is the range of gas saturation when gas relative permeability


is 10%?

krg
100% Gas

50%
40%

30%

20%

10%

5%

1%

100% Water 100% Oil

206
Question 9.6

Based on the available data below. Find the expected oil flow rate produced 18
ft above the free water level.

Parameter Value

µw 1 [cP]

µo 0.9 [cP]

rw 0.25 [ft]

D 5000 [ft]

k 200 [mD]

ф 0.2

Pwf 1200 [psig]

Pe 2200 [psig]

h 58 [ft]

Oil
0.8

0.6
kr

0.4

Water
0.2

0
0 0.2 0.4 0.6 0.8 1
Sw

35
Height above free water level [ft]

30

25

20

15

10

0
0 0.2 0.4 0.6 0.8 1
Sw

207
208
Chapter 10
Data Integration and
Volumetric Estimation of
Hydrocarbons
In this chapter, we will discuss how petrophysical data can be used to estimate
the hydrocarbons in a reservoir and beyond. Understanding the petrophysical
data helps in better understanding the reservoir with regard to where the
hydrocarbons reside, the amount of hydrocarbons available, and how effectively
the hydrocarbons can be extracted. This chapter covers some applications of the
petrophysical properties covered in Chapters 2–9 to estimate the hydrocarbons
in place with a few additional terms discussed. In addition, petrophysical data
and usage are also discussed in this chapter. Dealing with a reservoir can be
troublesome since it is underground and obscure; obtaining petrophysical data
from different sources and analyzing it makes the picture clearer.

10.1 Estimation of Hydrocarbons in Place

In Chapter 2, porosity was explained as a parameter that quantifies the storage


capacity of a reservoir to store fluids while fluid saturation, discussed in Chapter
5, quantifies the exact amount of fluids that reside in the empty spaces of a
reservoir. Oil initially in place (OIP) is the volume of oil in a reservoir, while gas
initially in place (GIP) is the gas volume in a reservoir.

In Chapter 5, oil saturation was defined as:

Vo
So = (10.1)
Vp

where So is the oil saturation [dimensionless], Vo is the volume of oil in a porous


medium [m3], and Vp is the pore volume of a porous medium [m3].

Rearranging the equation so that Vo becomes the subject yields:

Vo = So Vp (10.2)

209
In order to find the pore volume of a reservoir, the petroleum industry usually
uses seismic surveys to obtain the areal extent of the reservoir along with data
from pressure tests from well testing analysis. On the other hand, laboratory
measurements on extracted core samples along with wireline logging data can
give an estimate of the porosity in the reservoir. In Chapter 2, porosity was
defined as:

Vp
φ= (10.3)
Vb

where ф is the porosity [dimensionless] and Vb is the bulk volume of a porous


medium [m3].

Rearranging the equation so that Vp is the subject yields:

Vp = φVb (10.4)

Now substituting Equation 10.4 in Equation 10.2 yields:

Vo = So φVb (10.5)

Two notes can be considered:

1) Water saturation is usually measured by capillary pressure and


resistivity logs and thus the oil saturation is replaced by 1- Sw for an oil/
water and gas/water reservoirs.

2) Vo for the reservoir is referred to as OIP and thus the equation


becomes:

OIP = φVb (1 − Sw ) (10.6)

The same process would go for gas/water reservoir for GIP and the equation
becomes:

GIP = φVb (1 − Sw ) (10.7)

10.1.1 Net to Gross

There are other parameters that can be included in the OIP/GIP equations such
as the net to gross (NTG), which is the ratio of the fraction of reservoir section
to the total studied section (reservoir and non-reservoir sections). For example,
in Figure 10.1, the total section includes both reservoir rocks and impermeable
rocks. The fraction of the reservoir section to the total section in this case will be:

h1 + h2 + h3 + h4
NTG = (10.8)
htotal

210
NTG is a fraction that can range from 0 to 1: the higher the number, the higher
the reservoir fraction. Note that the location of the oil/water contact (OWC) can
affect the NTG value as reservoir rocks below the OWC (fully saturated with water)
are not considered to be viable and thus only reservoir rocks above the OWC are
considered. The NTG is often calculated using wireline logging tools.

If the NTG is known, then the OIP equation can be modified to obtain:

OIP = φ A htotal NTG (1 − Sw ) (10.9)

h1

h2

htotal
h3

h4

Impermeable
Reservoir Rock
rock

Figure 10.1: Schematic showing how to calculate the NTG from a reservoir with impermeable
rock layers.

10.1.2 Fluid Properties

The state of the fluids in the reservoir varies as a function of pressure and
temperature, as briefly discussed in Chapter 5. This means that the quantity of
oil in the reservoir will vary when it reaches the surface due to the change in
pressure and temperature; the same can be said about the gas in the reservoir.

Oil Reservoirs

The oil in the reservoir consists of a long chain of hydrocarbons when compared
to gas and contains dissolved gas. The initial reservoir pressure for oil reservoirs
is high and when the production starts, the reservoir pressure depletes. This
decrease in the reservoir pressure causes the dissolved gas in the oil to expand,
which in turn causes the total volume of oil to increase since the gas is in the
oil. The oil expansion will continue to increase until it reaches the bubble point
pressure where the first bubble of gas starts to leave the oil and the oil cannot
hold any more gas beyond this point. Any further reduction beyond the bubble
point pressure will cause more gas to leave the oil and thus lead to further
211
reduction in the oil volume. The oil formation volume factor (Bo) can be defined
as:
Volume of oil in the reservoir
Bo = (10.10)
Volume of oil on the surface
Figure 10.2 shows the relationship between Bo and pressure. Understanding Bo
is very important as the volume of oil in the reservoir will differ when it reaches
the surface. Based on the Bo relationship as a function of pressure (Figure 10.2),
the oil will shrink when it reaches the surface, making less oil available to be sold.

Since Bo is a volume divided by a volume, it is dimensionless. However, it is used


to convert the volume of oil from reservoir (res) conditions to surface conditions
(stock tank, st) and thus a unit of [res m3/st m3] is used. In this case, Bo will be
placed in the denominator of the OIP, Equation 10.6, and OIP will now become
STOIP, which is stock tank oil in place (N). Stock tank is the condition where the oil
is at surface conditions (low pressure and temperature) and hence the equation
will become:
φVb (1 − Sw )
N= (10.11)
Bo
The oil formation volume factor will always be greater than 1 as the denominator
in the Bo equation is the volume on the surface and that will always be less than
or at least equal to the volume of oil in the reservoir. This leads to a shrinkage
of oil volume on the surface. Note that the maximum Bo is at the bubble point
pressure as the dissolved gas in the oil expands to the maximum stage at that
point.

Bob

Boi
Bo

Pb Pi
P
Figure 10.2: Schematic showing the relationship between Bo and pressure, where Pb is the bubble
point pressure and Bob is the oil formation volume factor at the bubble point pressure which is
the maximum Bo value.

212
Gas Reservoirs

Gas behaves in a different manner than oil. When subjected to a high pressure,
gas becomes compressed; it expands as the pressure acting on it is reduced.
Similar to Bo, the gas formation volume factor (Bg) can be defined as:

Volume of gas in the reservoir


Bg = (10.12)
Volume of gas on the surface

In addition, it is used the same way in the GIP equation to become STGIP, which
is stock tank gas in place (G):

φVb (1 − Sw )
G= (10.13)
Bg
The relationship between Bg and pressure is shown in Figure 10.3. In this case,
the volume of gas will expand as it reaches the surface and the pressure should
always be below the bubble point pressure. The value of Bg is always less than 1
for the same reasons discussed before.

Bg

P
Figure 10.3 Schematic showing the relationship between Bg and pressure.

10.1.3 Layered Systems

Hydrocarbons-in-place calculations can be used in a layered reservoir shown in


Figure 10.4 with varying properties in each layer; the STOIP in this case will be:

n=3
φA 
N= hi (1 − Swi ) (10.14)
Bo i=1

213
This equation can be modified depending on the variables in the reservoir. All
the constants will be out of the sigma while the variables are in the sigma.

The equation can be customized depending on the available data. If the variables
increase as shown in Figure 10.5, then the STOIP becomes:

n=3
A 
N= hi φi (1 − Swi ) (10.15)
Bo i=1

ф SW1 h1
Layer 1

ф Layer 2 SW2 h2

ф Layer 3 SW3 h3

Figure 10.4: Schematic showing a layered reservoir with variable thickness and water saturation.

ф1 SW1 h1
Layer 1

ф2 Layer 2 SW2 h2

ф3 Layer 3 SW3 h3

Figure 10.5: Schematic showing a layered reservoir with variable thickness, porosity, and water
saturation.

214
Example 10.1
Find the OIP of a layered reservoir with the given data below. The reservoir
has a cylindrical shape with a constant diameter of 800 ft.

Layer ф [%] Sw [%] h [ft]

1 20 20 14

2 22 22 8

3 19 24 22

Solution
The area can be found using the equation of a cylinder:

A = πr2 = π4002 = 502655 ft2

Using Equation 10.15 of a composite system with constant area and


excluding Bo, we obtain:
n=3

OIP = A hi φi (1 − Swi )
i=1

Thus,

502655[14×0.2×(1−0.2)+8×0.22×(1−0.22)+22×0.19×(1−0.24)] = 3,412,826 ft3

10.1.4 Unit Systems

In the petroleum industry, oilfield units are the most commonly used units.
Therefore, a conversion needs to be made to convert to oilfield units. The oilfield
units are the following for each parameter: RB (reservoir barrel) is for OIP, STB
(stock tank barrel) is for STOIP, ft3 is for GIP, and SCF (standard cubic feet) is for
STGIP. In oilfield units, RB/STB is for Bo while ft3/SCF is for Bg.

The bulk volume is usually subdivided into the area (A) and thickness of the
hydrocarbon zone (h) and since the area is much larger than the thickness, it is
usually reported in acres while including the conversion factor from acres to ft2
in the equation for GIP and STGIP:

GIP = 43560φAh(1 − Sw ) (10.16)

215
where GIP is the gas in place (reservoir condition) [ft3], A is the area of the reservoir
[acres], h is the thickness of the reservoir [ft], ф is the porosity [dimensionless],
and Sw is the water saturation [dimensionless].

43560φAh(1 − Sw )
G= (10.17)
Bg
where G is the stock tank gas in place (STGIP, surface condition) [SCF], A is the
area of the reservoir [acres], h is the thickness of the reservoir [ft], ф is the
porosity [dimensionless], Sw is the water saturation [dimensionless], and Bg is the
gas formation volume factor [ft3/SCF].

For oil reservoirs, additional conversion needs to be made from ft3 to RB and
thus the OIP and STOIP will become:

OIP = 7758φAh(1 − Sw ) (10.18)

where OIP is the oil in place (reservoir condition) [RB], A is the area of the reservoir
[acres], h is the thickness of the reservoir [ft], ф is the porosity [dimensionless],
and Sw is the water saturation [dimensionless].

7758φAh(1 − Sw )
N= (10.19)
Bo
where N is the stock tank oil in place (STOIP, surface condition) [STB], A is the area
of the reservoir [acres], h is the thickness of the reservoir [ft], ф is the porosity
[dimensionless], Sw is the water saturation [dimensionless], and Bo is the oil
formation volume factor [RB/STB].

In addition, oilfield units use different prefixes, as shown in Table 10.1.

Table 10.1: Oilfield Prefixes

Prefix Symbol Multiplication Factor

Thousand M 103

Million MM 106

Billion MMM or B 109

Trillion T 1012

216
Example 10.2
Based on the information of a gas reservoir provided in the table below,
calculate the STGIP.

Parameter Value

Area of reservoir [acres] 1500

Thickness [ft] 100

Porosity [%] 10

Water saturation [%] 20

Gas formation volume factor [ft3/SCF] 0.0035

Solution
Initial gas in place is given by Equation 10.17, which yields:
43560φAh(1 − Sw ) 43560 × 1500 × 100 × 0.1 × (1 − 0.2)
G= = = 149.35 B SCF
Bg 0.0035

Example 10.3
Based on the information of an oil reservoir provided in the table below,
calculate the STOIP.

Parameter Value

Area of reservoir [acres] 1500

Thickness [ft] 100

Porosity [%] 10%

Water saturation [%] 20

Oil formation volume factor [RB/STB] 1.2

Solution
Initial oil in place is given by Equation 10.19, which yields:

7758φAh(1 − Sw ) 7758 × 1500 × 100 × 0.1 × (1 − 0.2)


N= = = 77.58 MM STB
Bo 1.2

217
10.2 Data Integration and Uncertainty

The petrophysical data studied in Chapters 2–9 can be used to find hydrocarbons
in place as well as to characterize and estimate the flow of hydrocarbons to the
surface.

Table 10.2 shows all the parameters in the STOIP and STGIP equations and the
tools to measure them.

Table 10.2: List of all the parameters used to calculate the hydrocarbons in place and the
tools used to find them.

Parameter Symbol Tool(s)

Porosity ф Porosity can be measured using wireline


porosity logging tools and laboratory
measurements on extracted core
samples.

Area A The areal extent of the reservoir can be


initially measured from seismic surveys
which then can be cross-checked by
pressure tests from well test analysis.

Thickness h The thickness of the hydrocarbon


column can be identified using several
tools such as resistivity logging tool,
capillary pressure, and RFT.

Water saturation Sw The water saturation can be measured


from extracted cores to the surface,
resistivity logs, and capillary pressure.
However, measuring the water
saturation of extracted cores is least
reliable due to the changes in the
phases of the fluids inside the cores
from the reservoir to the surface.

Oil and gas Bo The fluid formation volume factor can


formation and be measured using fluid sampling
volume factors Bg analysis that is different from
petrophysical analysis.

218
The other parameters discussed in Chapters 2–9 to characterize the flow behavior
and their usage are listed in Table 10.3.

Table 10.3: List of the petrophysical parameters used to characterize flow and their
usage.

Parameter Symbol Usage

Pore cp It is used to find the amount of


compressibility hydrocarbons produced due to
compaction that occurs as a result of
decrease in reservoir pressure due to
hydrocarbon production.

Absolute k It is used to estimate the single-phase


permeability flow rate of hydrocarbons from the
reservoir to the well. Single-phase
permeability is mainly used before
the injection of another fluid into the
reservoir occurs.

Wettability - It is mainly used to quantify the


effectiveness of oil recovery by water
injection.

Relative kr It is used to estimate the flow rate from


permeability the reservoir to the surface when more
than one phase is flowing such as in the
case of water injection.

Trying to estimate the amount of hydrocarbons is trying to solve a puzzle that


you cannot see. Having different tools to measure one parameter helps in
reducing the uncertainty of that parameter and makes the picture of the amount
of hydrocarbons underground somewhat clearer. Finding the exact value of the
hydrocarbons in place is impossible as we deal with a system that we do not
physically see. The petroleum industry is known to deal with a great uncertainty
all the time. Therefore, an industry standard is to report the hydrocarbons in
percentiles in which every parameter is measured by several tools if possible,
and statistical analysis is performed to identify an acceptable range of data for
the calculation. Since we now have a set of ranges for every parameter, multiple
outcomes will appear and the answer will not be unique, depending on the
parameters used. The range is usually inputted in a software and generated for
n number of runs. The higher the runs, the more time needed to find the output;
however, the results tend to be more reliable as it covers more options. After
obtaining the results, the software will order the results from the smallest to the
largest; at 10% of the data (low value) is P90, which is the 90th percentile, or a 90%

219
chance that the reservoir has this amount of hydrocarbon based on the input
information. As the value in the order becomes larger, the likelihood of obtaining
this value decreases. The three most common values reported in the petroleum
industry are P10, P50, and P90. Among these, P10 has the largest value (most
optimistic) but represents the least likelihood for the amount of hydrocarbons
present there. A common practice is to use P50 for further analysis.

Example 10.4
The data below are for an oil reservoir that composed of a single, fairly
homogeneous rock. The oil formation volume factor (Bo) is 1.31 bbl/STB
and water saturation in the oil zones is 20% based on resistivity logs and
capillary pressure data.

a) Based on the data below find the STOIP (N) using the
volumetric method.

b) What is the net to gross for this reservoir?

i. Reservoir Map (Area)

The top view of the reservoir area is shown below.

75 m
100 m

Oil Reservoir

ii. Saturation Height

RFT and height above the free water level are shown below.

220

P [psig]
5200

5250

5300
Depth [ft]

5350

5400

5450

5500

5550

25
Height above free water level [ft]

20

15

10

0
0 0.2 0.4 0.6 0.8 1
Sw

iii. Porosity and Lithology Map

5275 ft

5285 ft

5295 ft

5305 ft

5315 ft

5325 ft

5335 ft

5345 ft

5355 ft

5365 ft

5375 ft
Impermeable
Reservoir Rock
rock

221
Solution

a) First, we must calculate the area. The reservoir map shows 88 green
squares representing the reservoir, with the area of each square being
7,500 m2. Thus, the area of the reservoir is 66,000 m2 (or 7,104,181 ft2).

Then, we use the RFT plot to determine the free water level at 5,330 ft. This
follows that the oil water contact, OWC = FWL – 3 ft = 5,327 ft.

Using the calculated OWC, the oil zones exist at the following depths:

Zone 1 : 5,280 ft – 5295 ft


Zone 2 : 5,303 ft – 5,317 ft
Zone 3: 5,325 ft – 5,327 ft

Hence,
n=3

φhi = 0.2(15 + 14 + 2) = 6.2 ft
i=1

φAh(1 − Sw ) 7104181 × 6.2 × (1 − 0.2)


N= = = 4.79 MM STB
Bo 1.31 × 5.615

Note that a conversion factor is included to convert from ft3 to bbl.

b) N:G = 31/47 = 0.66.

10.3 Summary

One of the main goals for the petroleum industry is to estimate the amount of
hydrocarbons in place within a reservoir and then decide whether the reservoir
has commercial value to carry on production. Estimation of hydrocarbons in
place requires a proper understanding of petrophysical data in order to reduce
the associated uncertainty. The porosity and fluid saturation within the reservoir
can give an estimate of the amount of hydrocarbons in the reservoir. The fluid
properties should be analyzed since the fluids behave differently at different
levels beneath the ground due to changing temperatures and pressures. The
oil and gas formation volume factors are the ratios of the volume of oil/gas in
the reservoir to that on the surface and can be measured using fluid sampling
analysis. Since a great deal of uncertainty is involved, the hydrocarbons are
reported in percentiles, with 90th, 50th, and 10th percentiles being the most
commonly used ones.
222
End of Chapter Questions

Question 10.1

The green area in the map below shows the area of an oil reservoir and additional
information are shown below. Based on this information, find the OIP from the
data shown.

Top of the reservoir


0 ft

Layer 1
ф = 20%, Sw = 22%

Layer 2
ф = 18%, Sw = 24%

Layer 3
ф = 23%, Sw = 100%
120 ft

10 m
10 m

Question 10.2

A reservoir that has recently been discovered has three layers, with the top of the
reservoir marked at a depth of 4500 ft. Analyze both the rock data and fill in the
summary report attached.

223
10.2.1 – Fill in the blue boxes in the table below.

Core Depth D L A Vb Wdry Ws Vp Vm ф k


Label [ft] [cm] [cm] [cm2] [cm3] [g] [g] [cm3] [cm3] [%] [mD]

1 4503 3.79 7.63 11.28 86.08 183.54 202.21 18.34 67.74 6.0

2 4545 3.78 7.64 11.22 85.74 184.34 202.00 68.39 20.23 4.5

3 4550 3.76 7.62 11.10 84.61 185.01 201.58 16.28 68.33 19.24 4.0

4 4650 3.78 7.53 11.22 177.24 196.65 19.07 65.44 22.56 790.0

5 4700 3.78 7.66 11.22 85.96 178.84 19.43 66.53 22.60 810.7

6 4743 3.79 7.65 11.28 86.30 179.31 199.35 19.69 22.81 800.0

7 4805 3.77 7.62 11.16 85.06 189.00 24.12 60.94 28.35 230.0

8 4825 3.77 7.64 11.16 85.28 162.01 187.74 25.28 60.01 307.9

9 4847 3.76 7.65 11.10 84.94 163.49 188.63 300.0

Ws — is the weight of the core sample when it is 100% saturated with water
ρw — is 1.018 g/cc at experimental conditions

10.2.2 – Find the grain/matrix density for the following cores.

Core 1:

Core 4:

Core 9:

Zone Depth [ft]

1 4500 - 4645

2 4646 - 4750

3 4751 - 4850

224
10.2.3 – Find the rock compressibility for the following cores.

Rock compressibility [×10-6 1/psig]

Core 3

Core 6

Core 9

0.3

0.25

0.2
Core 3
Core 6
ΔVc p/Vp

0.15
Core 9

0.1

0.05

0
0 2000 4000 6000 8000 10000 12000
Applied Pressure (P)[psig]

10.2.4 – Fill in the blue boxes below for the single-phase brine permeability
data.

Core D L A q dP k k
Label [cm] [cm] [cm2] [cm3/s] [atm] [D] [mD]

2 3.78 7.64 11.22 0.005 0.8

5 3.78 7.66 11.22 0.015 0.8107 810.7

8 3.77 7.64 11.16 0.013 0.3079 307.9

μw — is 1.048 cP

225
10.2.5 – Find the equivalent liquid permeability from the gas permeability
data below.

1000

100
kg [mD]

Core 3
Core 6
10
Core 9

1
0 1 2 3 4 5
1/P [1/atm]

Core 3:

Core 6:

Core 9:

10.2.6 – Fill in the blue boxes in the table below.

Archie's Parameters a m n

Core from Zone 1 1.7 2

Core from Zone 2 1 1.9

Core from Zone 3 1 1.8

- The formation factor was measured on the core sample from zone 1
with a porosity of 19% to be 26.

- The formation factor was measured on the core sample from zone 2
with a porosity of 22.1% to be 17.6.

- The resistivity index as a function of water saturation of the core from


zone 3 is shown below:

226
100

10
Ir

Core from Zone 3

1
0.1 1
Sw

Depth [ft] фN [%] Rt [Ω.m] Sw [%]

4507 19.2 11.9 23.13

4520 18.9 11.5

4602 19.0 10.7 24.61

4655 22.0 15.5 15.41

4690 22.2 15.3

4722 22.9 14.8 15.17

4782 27.3 3.3 31.65

4818 27.6 3.1

Rw ­— is 0.025 Ω.m

Additional capillary pressure and height above the free water level data:

35 Core 2

30 Core 5
Core 7
25
Pc [psia]

20

15

10

0
0 0.2 0.4 0.6 0.8 1
Sw

227
350 Core 2

300 Core 5
Core 7
250

200
h [ft]
150

100

50

0
0 0.2 0.4 0.6 0.8 1
Sw

10.2.7 – Find the depth of the free water level (FWL) in feet based on the RFT
data below.

Repeat Formation Tester (RFT) Data

Pressure [MPa] Depth [m] Phase ρ [kg/m3]

16.20 1468 Oil 720

16.32 1483 Water 1015

10.2.8 – Find the contact angle shown below. The experiment is based on a
water droplet on core 3. Also, find the Amott index based on core 6.

4 mm

3 mm

10000

Core 6 - Water re-saturation


8000
Core 6 - Secondary drainage
6000

4000

2000
Pc [Pa]

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
-2000

-4000

-6000

-8000

-10000
Sw

228
10.2.9 – Fill in the blue boxes in the table below for the water relative
permeability data for core 8 and plot the data. Also, estimate the
displacement efficiency.

Sw So kw [mD] ko [mD] krw kro

0.325 0.675 0 274.0 0 0.89

0.373 0.627 0.77 227.8 0.0025 0.74

0.421 0.579 3.08 184.7 0.01 0.6

0.469 0.531 135.5 0.025 0.44

0.483 14.47 95.4 0.047 0.31

0.565 0.435 22.78 0.074 0.2

0.613 0.387 33.87 36.9 0.11

0.661 0.339 46.19 18.5 0.15 0.06

0.68 55.42 12.3 0.04

0.7 0.3 64.66 6.2 0.21 0.02

0.748 0.252 70.82 0 0.23 0

10.2.10 – Report Summary

1) Reservoir lithology:

2) Justification:

3) Average parameters

ф avg [%] Sw, avg [%]

Zone 1

Zone 2

Zone 3

4) Depth of the FWL [ft]:

5) Depth of the OWC [ft]:

6) Wettability of the reservoir:

7) Justification (three justifications required):

229
8) Estimate the STOIP using the average parameters and the oil
formation volume factor of 1.27 RB/STB. The area of the reservoir
found from the seismic survey is 1.1×106 ft2.

230
231
References

Ahmed, T. (2000). Reservoir engineering handbook. Houston, TX: Gulf Professional Publishing.

Alyafei, N. (2015). Capillary trapping and oil recovery in altered-wettability carbonate rock. PhD
thesis. Imperial College London.

Alyafei, N. and Blunt, M.J. (2016) . The effect of wettability on capillary trapping in carbonates.
Advances in Water Resources, 90, pp. 36–50.

Amyx, J., Bass, D. and Whiting, R. (1960). Petroleum reservoir engineering. New York: McGraw-
Hill.

Anderson, M.A. (2011). The Defining Series: Introduction To Wireline Logging. Schlumberger.

Blunt, M. J. (2017). Multiphase flow in permeable media: A pore-scale perspective. Cambridge


University Press.

Blunt, M.J. (2017). The Imperial College Lectures in Petroleum Engineering: Volume 2:
Reservoir Engineering. World Scientific.

BP. (2014). Enhanced oil recovery. London: BP

Dandekar, A. (2013). Petroleum reservoir rock and fluid properties. Boca Raton, FL: Taylor &
Francis.

Donaldson, E. and Alam, W. (2013). Wettability. Burlington: Elsevier Science.

Dullien, F. (2012). Porous media fluid transport and pore structure. Burlington:
Elsevier Science.

Masalmeh, S.K., (2002). The Effect of Wettability on Saturation Functions and Impact on Carbonate
Reservoirs in the Middle East. In: Abu Dhabi International Petroleum Exhibition and Conference. Society of
Petroleum Engineers.

Peters, E. (2012). Advanced petrophysics: Volume 1: Geology, porosity, absolute permeability,


heterogeneity, and geostatistics. Austin, TX: Live Oak Book Company.

Peters, E. (2012). Advanced petrophysics: Volume 2: Dispersion, interfacial phenomena/wettability,


capillarity/capillary pressure, relative permeability. Austin, TX: Live Oak Book Company.

Pinczewski, V. (2018). PTRL6001 Reservoir Engineering I. [online] Available at:


https://www.engineering.unsw.edu.au/petroleum-engineering/ptrl6001-reservoir-engineering-1
[Accessed 15 Aug. 2018].

Tiab, D. and Donaldson, E. (2008). Petrophysics. Boston: Gulf Professional Publishing.

Tran, N. H. (2018). PTRL3009 Reservoir Engineering B. [online] Available at:


https://www.engineering.unsw.edu.au/petroleum-engineering/ptrl3009-reservoir-engineering-B
[Accessed 15 Aug. 2018].

Von Gonten, W., McCain, W. and Wu, C. (2018). Course Notes for Petroleum Engineering
311 Reservoir Petrophysics. [online]. Available at: http://www.pe.tamu.edu/blasingame/
data/z_zCourse_Archive/P311_Reference/P311_Course_Notes_(pdf)/P311_1992C_Wu_Notes.pdf
[Accessed 15 Aug. 2018].

232
‘‘This excellent text book provides much-needed reference on reservoir rock properties.
Prof. Nayef Alyafei has based this work on his own popular lecture courses and his
extensive research in multiphase flow in porous media. A wide range of topics is presented
clearly with excellent illustrations and explanations throughout. The approach follows an easy-
to-follow coherent progression of ideas and is pedagogical in its presentation, making this work ideal
as a textbook for undergraduate or post-graduate studies in petroleum engineering, hydrology
or related disciplines. The book provides a much needed reference on petrophysics which is also
valuable for researchers and professionals working in the oil industry. It is also of interest to
the growing body of students, researchers, scientists and engineers working on flow in porous
media with a variety of applications from hydrocarbon recovery to carbon dioxide storage. I will
certainly recommend this work to my own students and colleagues, and use it in my teaching.”
Martin Blunt, Professor of Reservoir Engineering, Imperial College London

This book covers the essential concepts of rock properties aiding students, petroleum
geoscientists, and engineers to understand petroleum reservoirs.

Key Features:

• Explains the fundamental concepts with great clarity and a step-by-step approach.
• Provides numerous examples and problems on each covered topic.
• Written in clear English language to appeal to global students.
• Summary highlighting the main points of each chapter.
• Numerous illustrative figures to solidify the understanding of the concepts.

Dr. Nayef Alyafei received his Ph.D. in Petroleum Engineering from Imperial College London in 2015
and joined Texas A&M University as a faculty member in 2015. Dr. Alyafei is known for his passion in
teaching with a unique teaching style. He received several teaching awards such as Teaching Excellence Award
2019, Faculty of the year 2019, and the Distinguished Achievement College-level Award for Teaching. In
addition, he developed an undergraduate technical elective course and made significant curriculum changes to
several courses in the department. Dr. Alyafei's research focuses on multi-phase flow in porous media.

You might also like