Sunteți pe pagina 1din 143

Formation

Evaluation
Dr. Paul W.J. Glover

MSc Petroleum
Geology
Department of Geology
and Petroleum Geology
University of Aberdeen
UK

Contents
Copyright

Formation Evaluation MSc Course Notes

Contents
1.

Introduction

2.

Reservoir Fluids

3.

Reservoir Drives

19

4.

Coring, Preservation and Handling

33

5.

Porosity

43

6.

Single Phase Permeability

54

7.

Wettability

76

8.

Capillary Pressure

84

9.

Electrical Properties

95

10. Relative Permeability

104

11. Commissioning Studies

131

Abbreviations
References

Dr. Paul Glover

Page i

Formation Evaluation MSc Course Notes

Introduction

Chapter 1: Introduction
1.1

Introduction

This course aims to provide an understanding of the behaviour of fluids in reservoirs, and the
use of core analysis in the evaluation of reservoir potential. It is intended to give the end user
of special core analysis data an insight into the experimental techniques used to generate such
data and an indication of its validity when applied to reservoir assessment. It has been written
from the standpoint of a major oil industry operational support group, and is based upon the
substantial experience of working in such an environment.

1.2

Core Analysis and other Reservoir Engineering Data

Special core analysis (SCAL) is one of the main sources of data available to guide the
reservoir engineer in assessing the economic potential of a hydrocarbon accumulation. The
data sources can be divided into field and laboratory measurements as shown in Figure 1.1.
Laboratory data are used to support
field measurements which can be
subject to certain limitations, e.g.:
(i)

Fluid saturations may be


uncertain where actual formation
brine composition and resistivity
are not available.

(ii)

Permeability derived from well


test data may be reduced by
localised formation damage
(skin effects) and increased by
fractures.

Sedimentological data can be used to


predict areal and vertical trends in
rock properties and as an aid in the
correct choice of core for laboratory
measurements.
For core analysis to provide
meaningful data, due regard must be
given to the ways in which rock
properties can change both during the
coring procedure (downhole), core
preservation,
and
subsequent
laboratory treatment.

Dr. Paul Glover

Page 1

Formation Evaluation MSc Course Notes

Introduction

This report is intended as a guide to the reliability and usefulness of the various RCAL and
SCAL techniques generally available, and the ways which these techniques have, and will
continue to be, refined in the light of current research. Maximum benefit will only be
obtained from core analysis by full consultation between the reservoir engineer and the
laboratory core analyst; taking all available data into account.

1.3

Reservoir Fluids and Drives

Hydrocarbon reservoirs may contain any or all of three fluid phases. These are;

Aqueous fluids (brines),


Oils, and
Gases (hydrocarbon and non-hydrocarbon).

The distribution of these in a reservoir depends upon the reservoir conditions, the fluid
properties, and the rock properties. The fluid properties are of fundamental importance, and
will be studied in the first part of this course.
The natural energy of a reservoir can be used to facilitate the production of hydrocarbon and
non-hydrocarbon fluids from reservoirs. These sources of energy are called natural drive
mechanisms. However, there may still be producible oil in a reservoir when natural drive
mechanisms are exhausted. There exist artificial drive mechanisms that can then be used to
produce some of the remaining oil. The type of drive currently operating in a reservoir has a
strong control on the evaluation and management of the reservoir. Consequently, drive
mechanisms will also be reviewed as part of the course.

1.4

Routine Core Analysis (RCAL)

Routine core analysis attempts to give only the very basic properties of unpreserved core.
These are basic rock dimensions, core porosity, grain density, gas permeability, and water
saturation. Taken in context routine data can provide a useful guide to well and reservoir
performance, provided its limitations are appreciated. These limitations arise because routine
porosity and permeability measurements are always made with gases on cleaned, dried core at
room conditions. Such conditions are distinctly different from the actual reservoir situation.
Thus routine data should be applied to the reservoir state with caution. This is especially true
for permeability measurements. Routine core analysis data is cheap, and often form the great
majority of the dataset representing reservoir core data. A schematic diagram of common
RCAL measurements is given as Figure 1.2.
Routine porosity data are generally reliable, being little affected by interactions between
minerals and reservoir fluids. Correction for overburden loading is usually all that is required.
Routine permeability results can misrepresent the reservoir situation as reservoir fluids often
interact with the minerals forming the pore walls. This is frequently the case because these
interactions cannot be allowed for in routine measurements. Correction can be only made for
the compressibility of gases used. Thus the Klinkenberg correction converts gas permeability
to equivalent liquid permeability (KL) but still assumes no fluid-rock interaction. An actual
liquid, brine or oil, usually gives a lower permeability than KL. If interface sensitive clays are

Dr. Paul Glover

Page 2

Formation Evaluation MSc Course Notes

Introduction

present in the reservoir, drying can destroy them and KL may be one or two orders of
magnitude greater than an actual brine permeability measured on preserved, undried, core.
An example of this effect is seen in the Magnus field and was demonstrated by Heaviside,
Langley and Pallatt [1]. Permeability is affected by overburden loading to a greater extent than
porosity. This must be allowed for when applying routine data to the reservoir situation.

Each of the RCAL measurements made is discussed in detail, covering; the theory, test
methods, and limitations of alternative methods. The topics covered will include:
Chapter 4.
Chapter 5.
Chapter 6.

1.5

Unpreserved core cleaning and water analysis.


Sample dimension, porosity and grain density measurements.
Gas permeability.

Special Core Analysis (SCAL)

Special Core Analysis attempts to extend the data provided by routine measurements to
situations more representative of reservoir conditions. SCAL data is used to support log and
well test data in gaining an understanding of individual well and overall reservoir
performance. However, SCAL measurements are more expensive, and are commonly only
done on a small selected group of samples, or if a difficult strategic reservoir management
decision has to be made (e.g. to gasflood, or not to gasflood).
Tests are carried out to measure fluid distribution, electrical properties and fluid flow
characteristics in the two and occasionally three phase situation, and are made on preserved
core. A schematic diagram of common SCAL measurements is given as Figure 1.3.

Dr. Paul Glover

Page 3

Formation Evaluation MSc Course Notes

Introduction

Porosity and single phase gas or liquid permeabilities are measured at overburden loadings so
that the room condition data can be corrected.
Wettability and capillary pressure data are generated by controlled displacement of a wetting
phase by a non wetting phase e.g., brine by air, brine by oil or air by mercury. These systems
usually have known interfacial tension (IFT) and wetting (contact) angle properties.

Conversion to the required reservoir values of IFT and contact angle can then be attempted to
give data for predicting saturation at a given height within a reservoir. Electrical properties are
measured at formation brine saturations of unity and less than unity, to obtain the cementation
exponent, resistivity index, and excess conductivity of samples. These are used to provide
data for interpretation of down-hole logs.
Relative permeability attempts to provide data on the relative flow rates of phases present (e.g.
oil and water or gas and water). Fluid flow is strongly influenced by fluid viscosities, and
wetting characteristics. Care has to be taken that measurements are made under appropriate
conditions, which allow some understanding of the wetting characteristics. The data
generated allows relative flow rates and recovery efficiency to be assessed.
Each of the SCAL measurements made is discussed in detail in the relevant chapter, covering
the theory, test methods, and limitations of alternative methods. The topics covered will
include:

Dr. Paul Glover

Page 4

Formation Evaluation MSc Course Notes

Chapter 4.
Chapter 5.
Chapter 6.
Chapter 7.
Chapter 8.
Chapter 9.

Chapter 10.
Chapter 11.

1.6

Introduction

Preserved core; methods of preservation and requirement for


preserved core.
Porosity at overburden pressures.
Gas and liquid single phase permeabilities at overburden conditions.
Wettability determinations; techniques available and limitations of
data obtained.
Capillary pressure measurements; techniques available and
limitations of data obtained.
Electrical measurements; resistivity index and saturation exponent,
formation factor at room and overburden pressure, and cementation
exponent.
Relative Permeability; Theory, Techniques available, limitations
and application of data.
Typical SCAL programmes.

Arrangement of the Text

Effective assessment of reservoirs begins with an understanding of the properties of reservoir


fluids, which is covered in Chapter 2. Chapter 3 discusses the various reservoir drives
encountered in reservoir management. Chapter 4 discusses coring, core preservation and
handling, which is of relevance mainly to SCAL studies. Chapters 5 and 6 cover RCAL
porosity and permeability measurements, together with extensions to overburden pressure for
SCAL studies. Chapters 7 to 10 cover various wettability, capillary pressure, electrical, and
relative permeability measurements commonly practised in SCAL studies. Chapter 11 briefly
examines typical SCAL work programmes.

Dr. Paul Glover

Page 5

Formation Evaluation MSc Course Notes

Chapter 2:

Reservoir Fluids

Reservoir Fluids

2.1 Introduction
Reservoir fluids fall into three broad categories; (i) aqueous solutions with dissolved salts, (ii)
liquid hydrocarbons, and (iii) gases (hydrocarbon and non-hydrocarbon). In all cases their
compositions depend upon their source, history, and present thermodynamic conditions. Their
distribution within a given reservoir depends upon the thermodynamic conditions of the
reservoir as well as the petrophysical properties of the rocks and the physical and chemical
properties of the fluids themselves. This chapter briefly examines these reservoir fluid
properties.

2.2 Fluid Distribution


The distribution of a particular set of reservoir fluids depends not only on the characteristics
of the rock-fluid system now, but also the history of the fluids, and ultimately their source. A
list of factors affecting fluid distribution would be manifold. However, the most important
are:
Depth The difference in the density of the fluids results in their separation over time due to
gravity (differential buoyancy).
Fluid Composition The composition of the reservoir fluid has an extremely important
control on its pressure-volume-temperature properties, which define the relative volumes of
each fluid in a reservoir. This subject is a major theme of this chapter. It also affects
distribution through the wettability of the reservoir rocks (Chapter 7).
Reservoir Temperature Exerts a major control on the relative volumes of each fluid in a
reservoir.
Fluid Pressure Exerts a major control on the relative volumes of each fluid in a reservoir.
Fluid Migration Different fluids migrate in different ways depending on their density,
viscosity, and the wettability of the rock. The mode of migration helps define the distribution
of the fluids in the reservoir.
Trap-Type Clearly, the effectiveness of the hydrocarbon trap also has a control on fluid
distribution (e.g., cap rocks may be permeable to gas but not to oil).
Rock structure The microstructure of the rock can preferentially accept some fluids and not
others through the operation of wettability contrasts and capillary pressure. In addition, the
common heterogeneity of rock properties results in preferential fluid distributions throughout
the reservoir in all three spatial dimensions.
The fundamental forces that drive, stabilise, or limit fluid movement are:

Gravity (e.g. causing separation of gas, oil and water in the reservoir column)
Capillary (e.g. responsible for the retention of water in micro-porosity)
Molecular diffusion (e.g. small scale flow acting to homogenise fluid compositions within
a given phase)
Thermal convection (convective movement of all mobile fluids, especially gases)
Fluid pressure gradients (the major force operating during primary production)

Dr. Paul Glover

Page 6

Formation Evaluation MSc Course Notes

Reservoir Fluids

Although each of these forces and factors vary from reservoir to reservoir, and between
lithologies within a reservoir, certain forces are of seminal importance. For example, it is
gravity that ensures, that when all three basic fluids types are present in an
uncompartmentalised reservoir, the order of fluids with increasing depth is
GAS:OIL:WATER, in exact analogy to a bottle of french dressing that has been left to settle.

2.3 Aqueous Fluids


Accumulations of hydrocarbons are invariably associated with aqueous fluids (formation
waters), which may occur as extensive aquifers underlying or interdigitated with hydrocarbon
bearing layers, but always occur within the hydrocarbon bearing layers as connate water.
These fluids are commonly saline, with a wide range of compositions and concentrations;
Table 2.1 shows an example of a reservoir brine. Usually the most common dissolved salt is
NaCl, but many others occur in varying smaller quantities. The specific gravity of pure water
is defined as unity, and the specific gravity of formation waters increases with salinity at a rate
of about 0.075 per 100 parts per thousand of dissolved solids. When SCAL measurements are
made with brine, it is usual to make up a simulated formation brine to a recipe such as that
given in Table 2.1, and then deaerate it prior to use.
Table 2.1 Composition of Draugen 6407/9-4 Formation Water
Component

Concentration, g dm-3

Pure water
NaCl
CaCl2.6H2O
MgCl2.6H2O
KCl
NaHCO3
SrCl2.6H2O
BaCl2.6H2O

Solvent
34.70
4.90
2.70
0.40
0.40
0.12
0.06
Final pH = 7

Why a connate water phase is invariably present in hydrocarbon bearing reservoir rock is
easily explained. The reservoir rocks were initially fully or partially saturated with aqueous
fluids before the migration of the oil from source rocks below them. The oil migrates
upwards from the source rocks, driven by the differential buoyancy of the oil and the water. In
this process most of the water swaps places with the oil since no fluids can escape from the
cap rock above the reservoir. However, the water is not completely displaced as the initial
reservoir rock is invariably water-wet, leaving the water-wet grains covered in a thin layer of
water, with the remainder of the pore space full of oil. Water also remains in the microporosity where gravity segregation forces are insufficient to overcome the water-rock capillary
forces.
The aqueous fluids, whether as connate water or in aquifers, commonly contain dissolved
gases at reservoir temperatures and pressures. Different gases dissolve in aqueous fluids to
different extents, and this gas solubility also varies with temperature and pressure. Table 2.2
shows a selection of gases. If gas saturated water at reservoir pressure is subjected to lower

Dr. Paul Glover

Page 7

Formation Evaluation MSc Course Notes

Reservoir Fluids

pressures, the gas will be liberated, in exactly the same way that a lemonade bottle fizzes
when opened. In reservoirs the dissolved gas is mainly methane (from 10 SCF/STB at 1000
psi to 35 SCF/STB at 10 000 psi for gas-water systems, and slightly less for water-oil
systems). Higher salinity formation waters tend to contain less dissolved gas.
Table 2.2 Dissolution of Gases in Water (dissolved mole fraction) at 1 bar
104 Xgas @ 1 bar

Gas

Helium
Argon
Radon
Hydrogen
Nitrogen
Oxygen
Carbon dioxide
Methane
Ethane
Ammonium

25 C

55oC

0.06983
0.2516
1.675
0.1413
0.1173
0.2298
6.111
0.2507
0.3345
1876

0.07179
0.1760
0.8911
0.1313
0.08991
0.0164
3.235
0.1684
0.01896
1066

Xgas = mole fraction of gas dissolved at 1 bar pressure, i.e.=1/Hgas.


Aqueous fluids are relatively incompressible compared to oils, and extremely so compared to
gases (2.510-6 to 510-6 per psi decreasing with increasing salinity). Consequently, if a unit
volume of formation water with no dissolved gases at reservoir pressure conditions is
transported to surface pressure condition, it will expand only slightly compared to the same
initial volume of oil or gas. It should be noted that formation waters containing a significant
proportion of dissolved gases are more compressible than those that are not gas saturated.
These waters expand slightly more on being brought to the surface. However the reduction in
temperature on being brought to the surface causes the formation water to shrink and there is
also a certain shrinkage associated with the release of gas as pressure is lowered. The overall
result is that brines experience a slight shrinkage (< 5%) on being brought from reservoir
conditions to the surface.
Formation waters generally have densities that are greater than those of oils, and dynamic
viscosities that are a little lower (Table 2.3). The viscosity at high reservoir temperatures
(>250oC) can be as low as 0.3 cP, rises to above 1 cP at ambient conditions, and increases
with increasing salinity.

Dr. Paul Glover

Page 8

Formation Evaluation MSc Course Notes

Reservoir Fluids

Table 2.3 Densities and Viscosities for a Typical Formation Water and a Refined Oil
Brine Component

Composition, g/l

Pure water
NaCl
CaCl2.6H2O
MgCl2.6H2O
Na2SO4
NaHCO3

Solvent
150.16
101.32
13.97
0.55
0.21

Fluid

Temperature, oC

Density, g/cm3

Brine
Brine
Brine

20
25
30

1.1250
1.1237
1.1208

1.509
1.347
1.219

Kerosene
Kerosene
Kerosene

20
25
30

0.7957
0.7923
0.7886

1.830
1.661
1.514

Dynamic Viscosity, cP

2.4 Phase Behaviour of Hydrocarbon Systems


Figure 2.1 shows the pressure versus volume per mole weight (specific volume)
characteristics of a typical pure hydrocarbon (e.g. propane). Imagine in the following
discussion that all changes occur isothermally (with no heat flowing either into or out of the
fluid) and at the same temperature. Initially the component is in the liquid phase at 1000 psia,
and has a volume of about 2 ft3/lb.mol. (point A). Expansion of the system (AB) results in
large drops in pressure with small increases in specific volume, due to the small
compressibility of liquids (liquid hydrocarbons as well as liquid formation waters have small
compressibilities that are almost independent of pressure for the range of pressures
encountered in hydrocarbon reservoirs). On further expansion, a pressure will be attained
where the first tiny bubble of gas appears (point B). This is the bubble point or saturation
pressure for a given temperature. Further expansion (BC) now occurs at constant pressure
with more and more of the liquid turning into the gas phase until no more fluid remains. The
constant pressure at which this occurs is called the vapour pressure of the fluid at a given
temperature. Point C represents the situation where the last tiny drop of liquid turns into gas,
and is called the dew point. Further expansion now takes place in the vapour phase (CD).
The pistons in Figure 2.1 demonstrate the changes in fluid phase schematically. It is worth
noting that the process ABCD described above during expansion (reducing the
pressure on the piston) is perfectly reversible. If a system is in state D, then application of
pressure to the fluid by applying pressure to the pistons will result in changes following the
curve DCBA.

Dr. Paul Glover

Page 9

Formation Evaluation MSc Course Notes

Reservoir Fluids

We can examine the curve in Figure 2.1 for a range of fluid temperatures. If this is done, the
pressure-volume relationships obtained can be plotted on a pressure-volume diagram with the
bubble point and dew point locus also included (Figure 2.2). Note that the bubble point and
dew point curves join together at a point (shown by a dot in Figure 2.2). This is the critical
point. The region under the bubble point/dew point envelope is the region where the vapour
phase and liquid phase can coexist, and hence have an interface (the surface of a liquid drop or
of a vapour bubble). The region above this envelope represents the region where the
Dr. Paul Glover

Page 10

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Reservoir Fluids

Page 11

Formation Evaluation MSc Course Notes

Reservoir Fluids

vapour phase and liquid phase do not coexist. Thus at any given constant low fluid pressure,
reduction of fluid volume will involve the vapour condensing to a liquid via the two phase
region, where both liquid and vapour coexist. But at a given constant high fluid pressure
(higher than the critical point), a reduction of fluid volume will involve the vapour phase
turning into a liquid phase without any fluid interface being generated (i.e. the vapour
becomes denser and denser until it can be considered as a light liquid). Thus the critical point
can also be viewed as the point at which the properties of the liquid and the gas become
indistinguishable (i.e. the gas is so dense that it looks like a low density liquid and vice versa).
Suppose that we find the bubble points and dew points for a range of different temperatures,
and plot the data on a graph of pressure against temperature. Figure 2.3 shows such a plot.
Note that the dew point and bubble points are always the same for a pure component, so they
plot as a single line until the peak of Figure 2.2 is reached, which is the critical point.

The behaviour of a hydrocarbon fluid made up of many different hydrocarbon components


shows slightly different behaviour (Figure 2.4). The initial expansion of the liquid is similar to
that for the single component case. Once the bubble point is reached, further expansion does

Dr. Paul Glover

Page 12

Formation Evaluation MSc Course Notes

Reservoir Fluids

not occur at constant pressure but is accompanied by a decrease in pressure (vapour pressure)
due to changes in the relative fractional amounts of liquid to gas for each hydrocarbon in the
vaporising mixture. In this case the bubble points and dew points differ, and the resulting
pressure-temperature plot is no longer a straight line but a phase envelope composed of the
bubble point and dew point curves, which now meet at the critical point (Figure 2.5). There
are also two other points on this diagram that are of interest. The cricondenbar, which defines
the pressure above which the two phases cannot exist together whatever the temperature, and
the cricondentherm, which defines the temperature above which the two phases cannot exist
together whatever the pressure. A fluid that exists above the bubble point curve is classified
as undersaturated as it contains no free gas, while a fluid at the bubble point curve or below it
is classified as saturated, and contains free gas.
Figure 2.6 shows the PT diagram for a reservoir fluid, together with a production path from
the pressure and temperature existing in the reservoir to that existing in the separator at the

Dr. Paul Glover

Page 13

Formation Evaluation MSc Course Notes

Reservoir Fluids

surface. Note that the original fluid was an undersaturated liquid at reservoir conditions. On
production the fluid pressure drops fast with some temperature reduction occurring as the
fluid travels up the borehole. All reservoirs are predominantly isothermal because of their
large thermal inertia. This results in the production path of all hydrocarbons initially
undergoing a fluid pressure reduction. Figure 2.6 shows that the ratio of vapour to liquid at
separator conditions is approximately 55:45. If we analyse the PT characteristics of the
separator gas and separator fluid separately then we would find that the separator pressuretemperature point representing the separator conditions falls on the dew point line of the
separator gas PT diagram, and on the bubble point line of the separator oil PT diagram. This
indicates that the shape of the PT diagram for various mixtures of hydrocarbon gases and
liquids varies greatly. Clearly, therefore it is extremely important to understand the PT phase
envelope as it can be used to classify and understand major hydrocarbon reservoirs.

2.5 PVT Properties of Hydrocarbon Fluids


2.5.1 Cronquist Classification
Hydrocarbon reservoirs are usually classified into the following five main types, after
Cronquist, 1979:

Dry gas
Wet gas
Gas condensate
Volatile oil
Black oil

Dr. Paul Glover

Page 14

Formation Evaluation MSc Course Notes

Reservoir Fluids

Each of these
reservoirs can be
understood
in
terms of its phase
envelope.
The
typical
components
of
production from
each
of
these
reservoirs is shown
in Table 2.4, and a
schematic diagram
of their PT phase
envelopes
is
shown in Figure
2.7.

Table 2.4 Typical Mol% Compositions of Fluids Produced from Cronquist Reservoir
Types
Component or
Property
CO2
N2
C1
C2
C3
iC4
nC4
iC5
nC5
C6 s
C7+
GOR (SCF/STB)
OGR
(STB/MMSCF)
API Specific
Gravity, API
,oAPI
C7+ Specific
Gravity, o

Dry Gas

Wet Gas

Gas
Volatile Oil
Condensate

Black Oil

0.10
2.07
86.12
5.91
3.58
1.72
0.50
-

1.41
0.25
92.46
3.18
1.01
0.28
0.24
0.13
0.08
0.14
0.82

2.37
0.31
73.19
7.80
3.55
0.71
1.45
0.64
0.68
1.09
8.21

1.82
0.24
57.60
7.35
4.21
0.74
2.07
0.53
0.95
1.92
22.57

0.02
0.34
34.62
4.11
1.01
0.76
0.49
0.43
0.21
1.16
56.40

69000
15

5965
165

1465
680

320
3125

65.0

48.5

36.7

23.6

0.750

0.816

0.864

0.920

Note: Fundamental specific gravity o is equal to the density of the fluid divided by the
density of pure water, and that for C7+ is for the bulked C7+ fraction. The API specific gravity
API is defined as; API = (141.5/o) - 131.5.
Dr. Paul Glover

Page 15

Formation Evaluation MSc Course Notes

Reservoir Fluids

2.5.2 Dry Gas Reservoirs


A typical dry gas reservoir is shown in Figure 2.8. The reservoir temperature is well above
the cricondentherm. During production the fluids are reduced in temperature and pressure.
The temperature-pressure path followed during production does not penetrate the phase
envelope, resulting in the production of gas at the surface with no associated liquid phase.
Clearly, it would be possible to produce some liquids if the pressure is maintained at a higher
level. In practice, the stock tank pressures are usually high enough for some liquids to be
produced (Figure 2.9). Note the lack of C5+ components, and the predominance of methane in
the dry gas in Table 2.4.

2.5.3 Wet Gas Reservoirs


A typical wet gas reservoir is shown in Figure 2.9. The reservoir temperature is just above the
cricondentherm. During production the fluids are reduced in temperature and pressure. The
temperature-pressure path followed during production just penetrates the phase envelope,
resulting in the production of gas at the surface with a small associated liquid phase. Note the
presence of small amounts of C5+ components, and the continuing predominance of methane
in the wet gas in Table 2.4. The GOR (gas-oil ratio) has fallen as some liquid is being
produced. However, this liquid usually amounts to less than about 15 STB/MMSCF. Note
also the small specific gravity for C7+ components (0.750), indicating that the majority of the
C7+ fraction is made up of the lighter C7+ hydrocarbons.

Dr. Paul Glover

Page 16

Formation Evaluation MSc Course Notes

Reservoir Fluids

2.5.4 Gas Condensate Reservoirs


A typical gas condensate reservoir is shown in Figure 2.10. The reservoir temperature is such
that it falls between the temperature of the critical point and the cricondentherm. The
production path then has a complex history. Initially, the fluids are in an indeterminate vapour
phase, and the vapour expands as the pressure and temperature drop. This occurs until the
dewpoint line is reached, whereupon increasing amounts of liquids are condensed from the
vapour phase. If the pressures and temperatures reduce further, the condensed liquid may reevaporate, although sufficiently low pressures and temperatures may not be available for this
to happen. If this occurs, the
process is called isothermal
retrograde
condensation.
Isobaric
retrograde
condensation also exists as a
scientific phenomenon, but
does not occur in the
predominantly isothermal
conditions of hydrocarbon
reservoirs. Thus, in gas
condensate reservoirs, the
oil produced at the surface
results from a vapour
existing in the reservoir.
Note the increase in the C7+
components
and
the
continued importance of
methane in Table 2.4. The
GOR
has
decreased
significantly, the OGR has
increased, and the specific
gravity
of
the
C7+
components is increasing,
indicating
that
greater
fractions
of
denser
hydrocarbons are present in
the C7+ fraction.

2.5.5 Volatile Oil Reservoirs


A typical volatile oil reservoir is shown in Figure 2.11. The reservoir PT conditions place it
inside the phase envelope, with a liquid oil phase existing in equilibrium with a vapour phase
having gas condensate compositions. The production path results in small amounts of further
condensation, and re-evaporation can occur again, but should be avoided as much as possible
by keeping the stock tank pressure as high as possible. Reference to Table 2.4 shows that the
fraction of gases is reduced, and the fraction of denser liquid hydrocarbon liquids is increased,
compared with the previously discussed reservoir types. Changes in the GOR, OGR and
specific gravities are in agreement with the general trend.

Dr. Paul Glover

Page 17

Formation Evaluation MSc Course Notes

Reservoir Fluids

2.5.6 Black Oil Reservoirs

A typical gas condensate reservoir is shown in Figure 2.12. The reservoir temperature is
much lower than the temperature of the critical point of the system, and at pressures above the
cricondenbar. Thus, the hydrocarbon in the reservoir exists as a liquid at depth. The
production path first involves a reduction in pressure with only small amounts of expansion in
the liquid phase. Once the bubble point line is reached, gas begins to come out of solution and
continues to do so until the stock tank is reached. The composition of this gas changes very
little along the production path, is relatively lean, and is not usually of economic importance
when produced. Table 2.4 shows a produced hydrocarbon fluid that is now dominated by
heavy hydrocarbon liquids, with most of the produced gas present as methane. The GOR,
OGR and specific gravities mirror the fluid composition.

Dr. Paul Glover

Page 18

Formation Evaluation MSc Course Notes

Reservoir Drives

Chapter 3: Reservoir Drives


3.1 Introduction
Recovery of hydrocarbons from an oil reservoir is commonly recognised to occur in several
recovery stages. These are:
(i)
(ii)
(iii)
(iv)

Primary recovery
Secondary recovery
Tertiary recovery (Enhanced Oil Recovery, EOR)
Infill recovery

Primary recovery This is the recovery of hydrocarbons from the reservoir using the natural
energy of the reservoir as a drive.
Secondary recovery This is recovery aided or driven by the injection of water or gas from
the surface.
Tertiary recovery (EOR) There are a range of techniques broadly labelled Enhanced Oil
Recovery that are applied to reservoirs in order to improve flagging production.
Infill recovery Is carried out when recovery from the previous three phases have been
completed. It involves drilling cheap production holes between existing boreholes to ensure
that the whole reservoir has been fully depleted of its oil.
This chapter discusses primary, secondary and EOR drive mechanisms and techniques.

3.2 Primary Recovery Drive Mechanisms


During primary recovery the natural energy of the reservoir is used to transport hydrocarbons
towards and out of the production wells. There are several different energy sources, and each
gives rise to a drive mechanism. Early in the history of a reservoir the drive mechanism will
not be known. It is determined by analysis of production data (reservoir pressure and fluid
production ratios). The earliest possible determination of the drive mechanism is a primary
goal in the early life of the reservoir, as its knowledge can greatly improve the management
and recovery of reserves from the reservoir in its middle and later life.
There are five important drive mechanisms (or combinations). These are:
(i)
(ii)
(iii)
(iv)
(v)

Solution gas drive


Gas cap drive
Water drive
Gravity drainage
Combination or mixed drive

Table 3.1 shows the recovery ranges for each individual drive mechanism.

Dr. Paul Glover

Page 19

Formation Evaluation MSc Course Notes

Reservoir Drives

Table 3.1 Recovery ranges for each drive mechanism


Drive Mechanism
Solution gas drive

Energy Source

Recovery, % OOIP

Evolved solution gas and expansion

20-30

Evolved gas

18-25

Gas expansion

2-5

Gas cap drive

Gas cap expansion

20-40

Water drive

Aquifer expansion

20-60

Bottom

20-40

Edge

35-60

Gravity drainage

Gravity

50-70
A combination or mixed drive
occurs when any of the first
three drives operate together,
or when any of the first three
drives operate with the aid of
gravity drainage.
The reservoir pressure and
GOR trends for each of the
main (first) three drive
mechanisms is shown as
Figures 3.1 and 3.2. Note
particularly that water drive
maintains
the
reservoir
pressure much higher than the
gas drives, and has a uniformly
low GOR.

3.2.1 Solution Gas Drive


This drive mechanism requires
the reservoir rock to be
completely surrounded by
impermeable barriers.
As
production occurs the reservoir
pressure drops, and the
exsolution and expansion of
the dissolved gases in the oil
and water provide most of the
reservoirs drive energy. Small
amounts of additional energy
are also derived from the
expansion of the rock and
water, and gas exsolving and
Dr. Paul Glover

Page 20

Formation Evaluation MSc Course Notes

Reservoir Drives

expanding from the water phase. The process is shown schematically in Figure 3.3.
A solution gas drive reservoir is initially either considered to be undersaturated or saturated
depending on its pressure:

Undersaturated: Reservoir pressure > bubble point of oil.


Saturated: Reservoir pressure bubble point of oil.

For an undersaturated reservoir no free gas exists until the reservoir pressure falls below the
bubblepoint. In this regime reservoir drive energy is provided only by the bulk expansion of
the reservoir rock and liquids (water and oil).
For a saturated reservoir,
any oil production results in
a drop in reservoir pressure
that causes bubbles of gas
to exsolve and expand.
When the gas comes out of
solution the oil (and water)
shrink slightly. However,
the volume of the exsolved
gas, and its subsequent
expansion more than makes
up for this.
Thus gas
expansion is the primary
reservoir
drive
for
reservoirs below the bubble
point.
Solution
gas
drive
reservoirs show a particular
characteristic
pressure,
GOR and fluid production
history. If the reservoir is
initially undersaturated, the
reservoir pressure can drop
by a great deal (several
hundred psi over a few
months), see Figures 3.1
and 3.2.
This is because of the small
compressibilities of the
rock water and oil,
compared to that of gas. In
this undersaturated phase, gas is only exsolved from the fluids in the well bore, and
consequently the GOR is low and constant. When the reservoir reaches the bubble point
pressure, the pressure declines less quickly due to the formation of gas bubbles in the reservoir
that expand taking up the volume exited by produced oil and hence protecting against pressure
drops. When this happens, the GOR rises dramatically (up to 10 times). Further fall in
Dr. Paul Glover

Page 21

Formation Evaluation MSc Course Notes

Reservoir Drives

reservoir pressure, as production continues, can, however, lead to a decrease in GOR again
when reservoir pressures are such that the gas expands less in the borehole. When the GOR
initially rises, the oil production falls and artificial lift systems are then instituted.
Oil recovery from this type of reservoir is typically between 20% and 30% of original oil in
place (i.e. low). Of this only 0% to 5% of oil is recovered above the bubblepoint. There is
usually no production of water during oil recovery unless the reservoir pressure drops
sufficiently for the connate water to expand sufficiently to be mobile. Even in this scenario
little water is produced.

3.2.2 Gas Cap Drive


A gas cap drive reservoir usually benefits to some extent from solution gas drive, but derives
its main source of reservoir energy from the expansion of the gas cap already existing above
the reservoir.
The
presence
of
the
expanding gas cap limits the
pressure decrease experienced
by the reservoir during
production. The actual rate of
pressure decrease is related to
the size of the gas cap.
The GOR rises only slowly in
the early stages of production
from such a reservoir because
the pressure of the gas cap
prevents gas from coming out
of solution in the oil and
water.
As production
continues, the gas cap
expands pushing the gas-oil
contact (GOC) downwards
(Figure 3.4). Eventually the
GOC
will
reach
the
production wells and the
GOR will increase by large
amounts (Figures 3.1 and
3.2). The slower reduction in
pressure experienced by gas
cap reservoirs compared to
solution drive reservoirs
results in the oil production
rates being much higher
throughout the life of the
reservoir,
and
needing
artificial lift much later than
for solution drive reservoirs.
Dr. Paul Glover

Page 22

Formation Evaluation MSc Course Notes

Reservoir Drives

Gas cap reservoirs produce very little or no water.


The recovery of gas cap reservoirs is better than for solution drive reservoirs (20% to 40%
OOIP). The recovery efficiency depends on the size of the gas cap, which is a measure of
how much latent energy there is available to drive production, and how the reservoir is
managed, i.e. how the energy resource is used bearing in mind the geometric characteristics of
the reservoir, economics and equity considerations. Points of importance to bear in mind
when managing a gas cap reservoir are:
Steeply dipping reservoir oil columns are best.
Thick oil columns are best, and are perforated at the base, as far away from the gas cap as
possible. This is to maximise the time before gas breaks through in the well.
Wells with increasing GOR (gas cap breakthrough) can be shut in to reduce field wide
GOR.
Produced gas can be separated and immediately injected back into the gas cap to maintain
gas cap pressure.

3.2.3 Water Drive


The drive energy is provided by an aquifer that interfaces with the oil in the reservoir at the
oil-water contact (OWC). As production continues, and oil is extracted from the reservoir, the
aquifer expands into the reservoir displacing the oil. Clearly, for most reservoirs, solution gas
drive will also be taking place, and there may also be a gas cap contributing to the primary
recovery. Two types of water drive are commonly recognised:
Bottom water drive (Figure 3.5)
Edge water drive (Figure 3.5)
The pressure history of a water driven reservoir depends critically upon:
(i)
(ii)
(iii)

The size of the aquifer.


The permeability of the aquifer.
The reservoir production rate.

If the production rate is low, and the size and permeability of the aquifer is high, then the
reservoir pressure will remain high because all produced oil is replaced efficiently with water.
If the production rate is too high then the extracted oil may not be able to be replaced by water
in the same timescale, especially if the aquifer is small or low permeability. In this case the
reservoir pressure will fall (Figure 3.1).
The GOR remains very constant in a strongly water driven reservoir (Figure 3.2), as the
pressure decrease is small and constant, whereas if the pressure decrease is higher (weakly
water driven reservoir) the GOR increases due to gas exsolving from the oil and water in the
reservoir. Likewise the oil production from a strongly water driven reservoir remains fairly
constant until water breakthrough occurs.

Dr. Paul Glover

Page 23

Formation Evaluation MSc Course Notes

Reservoir Drives

Using analogous arguments to


the gas cap drive, it can be seen
that thick oil columns are again
an advantage, but the wells are
perforated high in the oil zone
to
delay
the
water
breakthrough.
When water
breakthrough does occur the
well can either be shut-down,
or assisted using gas lift.
Reinjection of water into the
aquifer is seldom done because
the injected water usually just
disappears into the aquifer with
no effect on aquifer pressure.
The recovery from water
driven reservoirs is usually
good (20-60% OOIP, Table
3.1), although the exact figure
depends on the strength of the
aquifer and the efficiency with
which the water displaces the
oil in the reservoir, which
depends on reservoir structure,
production well placing, oil
viscosity, and production rate.
If the ratio of water to oil
viscosity is large, or the
production rate is high then
fingering can occur which
leaves oil behind in the
reservoir (Figure 3.6).

Dr. Paul Glover

Page 24

Formation Evaluation MSc Course Notes

Reservoir Drives

3.2.4 Gravity Drainage


The density differences between oil and gas and water result in their natural segregation in the
reservoir. This process can be used as a drive mechanism, but is relatively weak, and in
practice is only used in combination with other drive mechanisms. Figure 3.7 shows
production by gravity drainage.

The best conditions for gravity drainage are:

Thick oil zones.


High vertical permeabilities.

The rate of production engendered by gravity drainage is very low compared with the other
drive mechanisms examined so far. However, it is extremely efficient over long periods and
can give rise to extremely high recoveries (50-70% OOIP, Table 3.1). Consequently, it is
often used in addition to the other drive mechanisms.

3.2.5 Combination or Mixed Drive


In practice a reservoir usually incorporates at least two main drive mechanisms. For example,
in the case shown in Figure 3.8. We have seen that the management of the reservoir for

Dr. Paul Glover

Page 25

Formation Evaluation MSc Course Notes

Reservoir Drives

different drive mechanisms


can be diametrically opposed
(e.g. low perforation for gas
cap reservoirs compared with
high perforation for water
drive reservoirs). If both
occur as in Figure 3.8, a
compromise must be sought,
and this compromise must
take into account the strength
of each drive present, the
size of the gas cap, and the
size/permeability of the
aquifer. It is the job of the
reservoir manager to identify
the strengths of the drives as
early as possible in the life of
the reservoir to optimise the
reservoir performance.

3.3 Secondary Recovery


Secondary recovery is the result of human intervention in the reservoir to improve recovery
when the natural drives have diminished to unreasonably low efficiencies. Two techniques
are commonly used:
(i)
(ii)

Waterflooding
Gasflooding

Dr. Paul Glover

Page 26

Formation Evaluation MSc Course Notes

Reservoir Drives

3.3.1 Waterflooding
This method involves the injection of water at the base of a reservoir to;
(i)
(ii)

Maintain the reservoir pressure, and


Displace oil (usually with gas and water) towards production wells.

The detailed treatment of waterflood recovery estimation, mathematical modelling, and design
are beyond the scope of these notes. However, it should be noted that the successful outcome
of a waterflood process depends on designs based on accurate relative permeability data in
both horizontal directions, on the choice of a good injector/producer array, and with full
account taken of the local crustal stress directions in the reservoir.

3.3.2 Gas Injection


This method is similar to waterflooding in principal, and is used to maintain gas cap pressure
even if oil displacement is not required. Again accurate relperms are needed in the design, as
well as injector/producer array geometry and crustal stresses. There is an additional
complication in that re-injected lean gas may strip light hydrocarbons from the liquid oil
phase. At first sight this may not seem a problem, as recombination in the stock tank or
afterwards may be carried out. However, equity agreements often give different percentages
of gas and oil to different companies. Then the decision whether to gasflood is not trivial (e.g.
Prudhoe Bay, Alaska).

3.4 Tertiary Recovery (Enhanced Oil Recovery)


Primary and secondary recovery methods usually only extract about 35% of the original oil in
place. Clearly it is extremely important to increase this figure. Many enhanced oil recovery
methods have been designed to do this, and a few will be reviewed here. They fall into three
broad categories; (i) thermal, (ii) chemical, and (iii) miscible gas. All are extremely
expensive, are only used when economical, and are implemented after extensive SCAL
studies have isolated the reservoir rock characteristics that are causing oil to remain
unproduced by conventional methods.

3.4.1 Thermal EOR


These processes use heat to improve oil recovery by reducing the viscosity of heavy oils and
vaporising lighter oils, and hence improving their mobility. The techniques include:
(i)
(ii)

Steam injection (Figure 3.9).


In situ combustion (injection of a hot gas that combusts with the oil in place, Figure
3.10).
(iii) Microwave heating downhole (3.11).
(iv) Hot water injection.
It is worth noting that the generation of large amounts of heat and the treatment of evolved gas
has large environmental implications for these methods. However, thermal EOR is probably
the most efficient EOR approach.

Dr. Paul Glover

Page 27

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Reservoir Drives

Page 28

Formation Evaluation MSc Course Notes

Reservoir Drives

3.4.2 Chemical EOR


These processes use chemicals added to water in the injected fluid of a waterflood to alter the
flood efficiency in such a way as to improve oil recovery. This can be done in many ways,
examples are listed below:
(i) Increasing water viscosity (polymer floods)
(ii) Decreasing the relative permeability to water (cross-linked polymer floods)
(iii) Increasing the relative permeability to oil (micellar and alkaline floods)

Dr. Paul Glover

Page 29

Formation Evaluation MSc Course Notes

Reservoir Drives

(iv) Decreasing Sor (micellar and alkaline floods)


(v) Decreasing the interfacial tension between the oil and water phases (micellar and
alkaline floods)
An example of chemical EOR is shown in Figure 3.12.

Dr. Paul Glover

Page 30

Formation Evaluation MSc Course Notes

Reservoir Drives

Chemical flood additives, especially surfactants designed to reduce surface or interfacial


tension, are extremely expensive. Thus the whole chemical EOR flood is designed to
minimise the amount of surfactants needed, and to ensure that the EOR process is
economically successful as well as technically. Chemical flooding is therefore not a simple
single stage process. Initially the reservoir is subjected to a preflush of chemicals designed to
improve the stability of the interface between the in-situ fluids and the chemical flood itself.
Then the chemical surfactant EOR flood is carried out. Commonly polymers are injected into
the reservoir after the chemical flood to ensure that a favourable mobility ratio is maintained.
A buffer to maintain polymer stability follows, then a driving fluid, which is usually water, is
injected. Figure 3.13 shows a typical flood sequence. Note that the mobilised oil bank moves
ahead of the surfactant flood, and how the total process has reduced the amount of the
surfactant fluid used.

3.4.3 Miscible Gas Flooding


This method uses a fluid that is miscible with the oil. Such a fluid has a zero interfacial
tension with the oil and can in principal flush out all of the oil remaining in place. In practice
a gas is used since gases have high mobilities and can easily enter all the pores in the rock
providing the gas is miscible in the oil. Three types of gas are commonly used:
(i) CO2
(ii) N2
(iii) Hydrocarbon gases.

Dr. Paul Glover

Page 31

Formation Evaluation MSc Course Notes

Reservoir Drives

All of these are relatively cheap to obtain either from the atmosphere or from evolved
reservoir gases. The high mobility of gases can cause a problem in the reservoir flooding
process, since gas breakthrough may be early due to fingering, leading to low sweep
efficiencies. Effort is then concentrated on trying to improve the sweep efficiency. One such
approach is called a miscible WAG (water alternating gas). In this approach water slugs and
CO2 slugs are alternately injected into the reservoir; the idea being that the water slugs will
lower the mobility of the CO2 and lead to a more piston-like displacement with higher flood
efficiencies. An additional important advantage of miscible gasflooding is that the gas
dissolves in the oil, and this process reduces the oil viscosity, giving it higher mobilities and
easier recovery. A WAG flood is shown in Figure 3.14.

3.5 Infill Recovery


Towards the end of the reservoir life (after primary, secondary and enhanced oil recovery), the
only thing that can be done to improve the production rate is to carry out infill drilling,
directly accessing oil that may have been left unproduced by all the previous natural and
artificial drive mechanisms. Infill drilling can involve very significant drilling costs, while the
resulting additional production may not be great.

Dr. Paul Glover

Page 32

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

Chapter 4: Coring, Preservation and Handling


4.1 Introduction
Large financial resources are invested in RCAL and SCAL core analysis programmes, and a
wide range of accurate experimental determinations can be carried out. However, cores are
expensive to obtain and represent a very dilute sampling of the reservoir rock. It is clear that
the samples used in such studies should be as representative as possible of the reservoir rock
at depth if the final data is to be credible, and an efficient use of the financial resources
devoted to them. Samples of the reservoir rock and the fluids they contain can be, and are
commonly, altered by the process of obtaining them (coring, recovery, wellsite handling,
shipment, storage, and preparation for experimentation). This chapter gives an overview of the
alteration processes that may be at work, together with some of the techniques available to
reduce alteration, and preserve the rock and fluid properties. The choice of core preparation
techniques is increasingly being made by using pre-screening information on the preserved
core. This approach is highly recommended.

4.2 The Coring Process


Reservoir rock undergoes changes during the coring process and on storage before reaching
the laboratory. The changes which occur are shown in Figure 4.1. Some of the changes are
reversible whilst others are irreversible but preventable. In most cases it is possible to leave
all or part of the core in a usable state. It is essential to use preserved core for certain SCAL
tests and for meaningful assessment of routine data.

Dr. Paul Glover

Page 33

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

Drilling of the core is invariably carried out at very high bottom hole pressure differentials,
thus the core is effectively water-flooded with mud filtrate, and the original contents partly
displaced. The outer surface of the core will be invaded by mud particles; the depth of
invasion being dependent upon permeability. This zone should be avoided when sampling.
The rest of the core will have had its original hydrocarbon content, and formation water
displaced by mud filtrate; the extent depending upon the core permeability and original fluid
saturations. These changes are not always harmful as the core can usually be restored in the
laboratory. More important changes can occur if the rock contains minerals sensitive to water
salinity. For example, contact with low salinity water can mobilise poorly adhered clay
particles, giving a small possibility that core can arrive in the laboratory with mobilised fines,
which are not significantly mobile in the reservoir. In a similar fashion the wetting
characteristics of the rock may be altered by surfactant mud additives. These changes are
usually unavoidable but if formations are known to be particularly sensitive, it may be
possible to modify mud composition and reduce overpressure to minimise damage. For
complete preservation of wettability on cores above the transition zone, coring with lease
crude is necessary. Water saturation may then also be retained intact, allowing better
estimation of initial reservoir oil saturation. For transition and water zone a bland mud
formulation will do the least harm to original rock properties.

Drying can be the worst that can happen to core after removal from the barrel. If interface
sensitive clays, e.g., fibrous illite are present they can be irreparably damaged by drying
(Figure 4.2) and any permeability measurements made on such core will be valueless. Thus it
Dr. Paul Glover

Page 34

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

is necessary to preserve some core in the state that it leaves the barrel either by immersion in
simulated formation brine or by wrapping in foil and wax. The latter technique is the
minimum required for samples intended for wettability measurements, but for straightforward
assessment of water zone permeabilities immersion in brine is adequate. The necessity for
preserved core will be more fully covered under relevant sections below.

4.3 Plug Sampling and Cleaning (Unpreserved Core)


Standard techniques are applied unless the core is very heterogeneous or likely to be damaged
by routine cleaning methods.
One or one and a half inch diameter
sample plugs are drilled and
trimmed to between two and three
inches long with simulated
formation brine as lubricant. If the
composition of formation brine is
unknown, a five percent sodium
chloride brine is used. Plugs are
taken at regular intervals (often
every 25 cm), parallel to bedding
planes for horizontal permeability
(see Figure 4.3a). Further plugs
normal to the bedding plane are
taken if required for vertical
permeability.
The sampling
interval can either, be increased, if
the core is from a formation known
to be homogeneous; or varied if the
core contains thin shaly bands
making it difficult to produce intact
plugs.
Thin shaly bands are
avoided unless frequent and
representative. Figure 4.3b analyses
the suitability of core plugs for
homogeneous, thickly bedded and
thinly bedded whole core.
Tests may also be carried out on full diameter core samples. This is necessary if plug sized
samples do not contain a representative pore size spectrum. Fractures, vugs (very large pores)
and stylolytes are typical structural features which necessitate measurement on full diameter
(whole core) samples. The measurements made are the same as for plug samples, but a
special core holder is necessary if horizontal permeabilities are required.
Plugs are cleaned by alternate extraction with hot toluene and methanol in Soxhlet extractors
(Figure 4.4a and 4.4b) until no further discolouration of solvent occurs. This may take from a
few, to several hundred hours depending upon permeability. Low permeability plugs are
seldom completely free of residual brine and oil at this stage. Complete removal of residual
fluids can only be achieved by prolonged Soxhlet extraction. Cores can also be cleaned by
Dr. Paul Glover

Page 35

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

flushing the core with alternate miscible solvents (e.g. toluene (for the oil phase) and
methanol (for the water phase)) done hot or cold in a Hassler coreholder (Figure 4.4a; also see
section 4.5). Both the aqueous (methanol) and oleic (toluene) cleaning phases exiting the rock
can be bulked and submitted for analysis of the amount of water and individual hydrocarbons
present.

Dr. Paul Glover

Page 36

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

Plugs are then dried to constant weight in a humidity controlled oven at 60C, 40% relative
humidity. Humidity controlled drying assists in restoring clays to nearer their reservoir state,
and may assist in preventing any further damage. However, the Klinkenberg corrected
equivalent liquid permeability from this type of drying process may still be larger than the
actual brine permeability due to the destruction of the clay texture.
If samples of plugs containing clays that are sensitive to drying are required for SEM analysis
(e.g. Figure 4.2), then a sample of the core with the original fluid contents must be critical
point dried. Ordinary drying destroys fine clay minerals because the interfacial forces
associated with the retreating liquid-vapour interface are high enough to mash the clay
structure. Critical point drying involves keeping a small sample of the core at pressure and
temperature conditions of the critical point of the fluids. The fluids will then be evaporated
from the sample without a liquid-vapour interface, which avoids destroying the fine clay
structure. This is an expensive operation because it can take many days to perform on even the
smallest sample chip. Consequently, it is almost never carried out for core plugs.

Dr. Paul Glover

Page 37

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Coring, Preservation and Handling

Page 38

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

4.4 Core and Plug Preservation (SCAL Techniques)


Preserved core is almost always required for one or more of the following reasons:
(i)
(ii)
(iii)
(iv)
(v)

Wettability determinations.
Prevention of drying of interface sensitive clays.
Maintenance of fluid saturations as received at surface.
Other SCAL where drying is not desirable.
Unconsolidated or relatively uncompacted samples that exhibit strong porosity and
permeability reductions with overburden stress.

Several methods of preservation are currently available and a choice can be made if the
requirement for preserved core is specified. The methods are:
Under simulated formation brine or kerosene, for water and oil zone cores respectively.
Cores are either kept under simulated formation brine in polymer containers with an airtight
seal at ambient pressure (certain types of spaghetti jars are good for this); see Figure 4.5.
Wax coated, for all SCAL purposes and especially wettability and residual oil
saturations. This technique, also called seal-peel, is widely used, and involves wrapping the
core in layers of plastic and aluminium foil before being dipped in wax. Cores preserved in
this way at the well site can be safely stored for moderately long periods and then be used for
almost all SCAL purposes (Figure 4.5).
In deoxygenated formation
brine or kerosene, for
wettability
measurements.
Samples are kept in anaerobic
jars which can be pressurised
to 30 psi (Figure 4.5). The
freshly cut core pieces are
placed in the jars under
deaerated simulated formation
brine or kerosene, and the jars
are then sealed. The remaining
air is then purged with
nitrogen, which is then raised
to 30 psi pressure. The samples
are then preserved under
reservoir fluid and a blanket of
inert gas. Providing that the
pressure is maintained, the
samples may be stored in this
state for long periods.
Wrapped in cling film and
frozen in solid CO2 for fluid
saturation
measurements.
This is used for unconsolidated
Dr. Paul Glover

Page 39

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

core. The samples are cooled using liquid nitrogen and are loaded into special containers. The
containers can be transported packed in solid CO2, and stored in special freezers. Plugs can be
cut from the core using liquid nitrogen as the cutting fluid, and the plugs are then immediately
loaded into special coreholders again, and stored frozen. The sample is thawed out and tested
without being removed from the special coreholders in which they were initially loaded.

4.5 Cleaning and Treatment of Preserved Core


Treatment of preserved core for the tests mentioned above will be reviewed with the
appropriate tests; but in general, sample plugs are drilled and trimmed using deoxygenated
formation brine and stored under deoxygenated, depolarised kerosene or brine before testing.
There are several methods of cleaning core. The actual method used will depend upon the
properties of the core. Usually the optimum method will be clear from pre-screening
information on the core. Pre-screening measurements include:

Core description
Core lithology
Assessment of consolidation
SEM analysis of mineralogy and pore structure
Petrographic analysis of mineralogy and pore structure
XRD/XRF analysis for bulk and clay mineralogies
CT scanning to assess core heterogeneities, Figure 4.6 (cross-bedding, and fractures)

Dr. Paul Glover

Page 40

Formation Evaluation MSc Course Notes

Coring, Preservation and Handling

This information is designed to identify possible problems with; (i) unconsolidated core, (ii)
clay sensitivity, (iii) stress sensitivity, (iv) core mineralogical hetereogeneity, (v) core
structural heterogeneity (e.g. fractures, vugs, fossils, and cross-bedding).
The commoner specialist cleaning methods include:
(i)
(ii)
(iii)
(iv)

Critical point drying


Cold miscible solvent flushing
Hot miscible solvent flushing
Direct fluid replacement (oil for oil and brine for brine)

Core cleaning, where appropriate, is most often carried out using miscible solvent flushing
techniques. The core if confined in a Hassler holder (Figure 4.4a) and cold solvent flowed
through it. Cleaning is usually complete after flowing three 200 ml alternating portions each
of methanol and toluene. Under certain circumstances only one portion of each solvent will
be used, although it is commoner to use at least three portions of each. This is applied to
cores known to contain mobile fines or where it is necessary to retain wettability modifying
crude oil components in their existing state. In some circumstances the evolved solvents need
to be quantitatively tested using chemical techniques for the water content, and the oil content
and composition. In this case special dry methanol is used, and the toluene is replaced with a
more efficient solvent such as CS2 (very dangerous) or dichloromethane.

4.6 Unconsolidated Core


Unconsolidated core gives rise to particular problems in coring, storage, handling and
plugging. Its extremely friable nature means that any rough handling damages the pore
structure irreversibly, and samples can turn into a pile of mud in your hand. The most
common method of handling, shipping, storage, and plugging this type of core is in a frozen
state. The core is frozen with liquid nitrogen or dry ice as soon as it emerges from the coring
barrel. It is then placed in a special core holder for the relevant experiment to be carried out.
Thawing inside the coreholder, prior to the experiment is only carried out after the sample has
been fully supported with the relevant applied confining pressures (see above).

4.7 Water Analysis


It is possible to obtain the initial water saturation and water composition from preserved
whole core and core plugs by extracting the water. This is done by the Dean and Stark
method. Figure 4.7 shows the Dean and Stark apparatus. The preserved sample is placed in a
paper thimble in the large glass container and fluxed with hot solvent. The water evaporates,
is carried by the solvent vapours into the long straight condenser in the top of the apparatus,
cools, condenses and is trapped in the graduated part of the apparatus. The water saturation
can be calculated by using the volume of the evolved water and a measurement of the porosity
of the rock sample after the extraction process. The composition of the evolved fluids can also
be analysed chemically, however, the water compositions more commonly used in SCAL
applications derive from wireline formation testing.

Dr. Paul Glover

Page 41

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Coring, Preservation and Handling

Page 42

Formation Evaluation MSc Course Notes

Porosity

Chapter 5: Porosity
5.1 Introduction and Definition
Total porosity is defined as the fraction of the bulk rock volume V that is not occupied by
solid matter. If the volume of solids is denoted by Vs, and the pore volume as Vp = V - Vs, we
can write the porosity as:

V - Vs Vp
Pore Volume
=
=
=
V
V Total Bulk Volume

(5.1)

The porosity can be expressed either as a fraction or as a percentage. Two out of the three
terms are required to calculate porosity.
It should be noted that the porosity does not give any information concerning pore sizes, their
distribution, and their degree of connectivity. Thus, rocks of the same porosity can have
widely different physical properties. An example of this might be a carbonate rock and a
sandstone. Each could have a porosity of 0.2, but carbonate pores are often very unconnected
resulting in its permeability being much lower than that of the sandstone.
A range of differently defined porosities are recognised and used within the hydrocarbon
industry. For rocks these are:
(i)
(ii)
(iii)
(iv)

Total porosity
Connected porosity
Effective porosity
Primary porosity

(v) Secondary porosity


(vi) Microporosity
(vii) Intergranular porosity
(viii) Intragranular porosity
(ix) Dissolution porosity
(x) Fracture porosity
(xi) Intercrystal porosity
(xii) Moldic porosity
(xiii) Fenestral porosity
(xiv) Vug porosity

Defined above.
The ratio of the connected pore volume to the total volume.
The same as the connected porosity.
The porosity of the rock resulting from its original depositional
structure.
The porosity resulting from diagenesis.
The porosity resident in small pores (< 2 m) commonly
associated with detrital and authigenic clays.
The porosity due to pore volume between the rock grains.
The porosity due to voids within the rock grains.
The porosity resulting from dissolution of rock grains.
The porosity resulting from fractures in the rock at all scales.
Microporosity existing along intercrystalline boundaries usually
in carbonate rocks.
A type of dissolution porosity in carbonate rocks resulting in
molds of original grains or fossil remains.
A holey (birds-eye) porosity in carbonate rocks usually
associated with algal mats.
Porosity associated with vugs, commonly in carbonate rocks.

It should be noted that if the bulk volume and dry weight, or the bulk volume, saturated
weight and porosity of a rock sample is known, then the grain density can be calculated. This
parameter is commonly calculated from the data to compare the results with the known grain

Dr. Paul Glover

Page 43

Formation Evaluation MSc Course Notes

Porosity

densities of minerals as a QA check. For example the density of quartz is 2.65 g cm-3, and a
clean sandstone should have a mean grain density close to this value.

5.2 Controls on Porosity


The initial (pre-diagenesis) porosity is affected by three major microstructural parameters.
These are grain size, grain packing, particle shape, and the distribution of grain sizes.
However, the initial porosity is rarely that found in real rocks, as these have subsequently been
affected by secondary controls on porosity such as compaction and geochemical diagenetic
processes. This section briefly reviews these controls.

5.2.1 Grain Size


The equilibrium porosity of a
porous material composed of a
random packing of spherical
grains is dependent upon the
stability given to the rock by
frictional and cohesive forces
operating between individual
grains.
These
forces
are
proportional to the exposed
surface area of the grains. The
specific surface area (exposed
grain surface area per unit solid
volume) is inversely proportional
to grain size. This indicates that,
when all other factors are equal, a
given weight of coarse grains will
be stabilised at a lower porosity
than the same weight of finer
grains. For a sedimentary rock
composed of a given single grain
size this general rule is borne out
in Figure 5.1 (to the left). It can
be seen that the increase in porosity only becomes significant at grain sizes lower than 100
m, and for some recent sediments porosities up to 0.8 have been measured. As grain size
increases past 100 m, the frictional forces decrease and the porosity decreases until a limit is
reached that represents random frictionless packing, which occurs at 0.399 porosity, and is
independent of grain size. No further loss of porosity is possible for randomly packed spheres,
unless the grains undergo irreversible deformation due to dissolution-recrystallisation,
fracture, or plastic flow, and all such decreases in porosity are termed compaction.

5.2.2

Grain Packing

The theoretical porosities for various grain packing arrangements can be calculated. The
theoretical maximum porosity for a cubic packed rock made of spherical grains of a uniform

Dr. Paul Glover

Page 44

Formation Evaluation MSc Course Notes

Porosity

size is 0.476, and is independent of grain size. The maximum porosity of other packing
arrangements is shown in Table 5.1 and Figure 5.2.
Table 5.1 Maximum porosity for different packing arrangements
Packing

Maximum Porosity (fractional)

Random
Cubic
Orthorhombic
Rhombohedral
Tetragonal

0.399 (dependent on grain size)


0.476
0.395
0.260
0.302

Figure 5.2 The porosities of standard packing arrangements.

5.2.3

Grain Shape

This parameter is not widely understood. Several studies have been carried out on random
packings of non-spherical grains, and in all cases the resulting porosities are larger than those
for spheres. Table 5.2 shows data for various shapes, where the porosity is for the frictionless
limit. Figure 5.1 shows data comparing rounded and angular grains, again showing that the
porosity for more angular grains is larger than those that are sub-spherical.
Table 5.2 The effect of grain shape on porosity
Grain Shape

Maximum Porosity (fractional)

Sphere
Cube
Cylinder
Disk

0.399 (dependent on grain size)


0.425
0.429
0.453

Dr. Paul Glover

Page 45

Formation Evaluation MSc Course Notes

5.2.4

Porosity

Grain Size Distribution

Real rocks contain a distribution of grain sizes, and often the grain size distribution is multimodal. The best way of understanding the effect is to consider the variable admixture of
grains of two sizes (Figure 5.3).

Figure 5.3 The behaviour of mixing grain sizes. Note that a mixture of two sizes has
porosities less than either pure phase.
The porosity of the mixture of grain sizes is reduced below that for 100% of each size. There
are two mechanisms at work here. First imagine a rock with two grain sizes, one of which has
1/100th the diameter of the other. The first mechanism applies when there are sufficient of the
larger grains to make up the broad skeleton of the rock matrix. Here, the addition of the
smaller particles reduces the porosity of the rock because they can fit into the interstices
between the larger particles. The second mechanism is valid when the broad skeleton of the
rock matrix is composed of the smaller grains. There small grains will have a pore space
between them. Clearly, if some volume of these grains are removed and replaced with a single
solid larger grain, the porosity will be reduced because both the small grains and their
associated porosity have been replaced with solid material. The solid lines GR and RF or RM
in Figure 5.3 represent the theoretical curves for both processes. Note that as the disparity
between the grain sizes increases from 6:3 to 50:5 the actual porosity approaches the
theoretical lines. Note also that the position of the minimum porosity is not sensitive to the
grain diameter ratio. This minimum occurs at approximately 20 to 30% of the smaller particle
diameter. In real rocks we have a continuous spectrum of grain sizes, and these can give rise
to a complex scenario, where fractal concepts become useful.

5.2.5 Secondary Controls on Porosity


Porosity is also controlled by a huge range of secondary processes that result in compaction
and dilatation. These can be categorised into (i) mechanical processes, such as stress

Dr. Paul Glover

Page 46

Formation Evaluation MSc Course Notes

Porosity

compaction, plastic deformation, brittle deformation, fracture evolution etc., and (ii)
geochemical processes, such as dissolution, repreciptation, volume reductions concomitant
upon mineralogical changes etc. The effect of stress mediated compaction on porosity will be
discussed in section 5.4. The effect of chemical diagenesis is more complex, and is better
assessed for any given rock by examination of SEM or optical photomicrographs.

5.3 Laboratory Determinations


There are many methods for measuring porosity, a few of which will be discussed below.
Several standard techniques are used. In themselves these are basic physical measurements of
weight, length, and pressures. The precision with which these can be made on plugs is
affected by the nature (particularly surface texture) of the plugs.

5.3.1

Direct Measurement

Here the two volumes V and Vs are determined directly and used in Eq. (1). This method
measures the total porosity, but is rarely used on rocks because Vs can only be measured if
the rock is totally disaggregated, and cannot, therefore, be used in any further petrophysical
studies. This measurement is the closest laboratory measurement to density log derived
porosities.

5.3.2

Imbibition Method

The rock sample is immersed in a wetting fluid until it is fully saturated. The sample is
weighed before and after the imbibition, and if the density of the fluid is known, then the
difference in weight is Vp , and the pore volume Vp can be calculated. The bulk volume V
is measured using either vernier callipers and assuming that the sample is perfectly cylindrical,
or by Archimedes Method (discussed later), or by fluid displacement using the saturated
sample. Vp and V can then be used to calculate the connected porosity. This is an accurate
method, that leaves the sample fully saturated and ready for further petrophysical tests. The
time required for saturation depends upon the rock permeability.

5.3.3

Mercury Injection

The rock is evacuated, and then immersed in mercury. At laboratory pressures mercury will
not enter the pores of most rocks. The displacement of the mercury can therefore be used to
calculate the bulk volume of the rock. The pressure on the mercury is then raised in a stepwise
fashion, forcing the mercury into the pores of the rock (Figure 5.4). If the pressure is
sufficiently high, the mercury will invade all the pores. A measurement of the amount of
mercury lost into the rock provides the pore volume directly. The porosity can then be
calculated from the bulk volume and the pore volume. Clearly this method also measures the
connected porosity. In practice there is always a small pore volume that is not accessed by the
mercury even at the highest pressures. This is pore volume that is in the form of the minutest
pores. So the mercury injection method will give a lower porosity than the two methods
described above. This is a moderately accurate method that has the advantage that it can be
done on small irregular samples of rock, and the disadvantage that the sample must be
disposed of safely after the test.

Dr. Paul Glover

Page 47

Formation Evaluation MSc Course Notes

Porosity

The mercury method also has the advantage that the grain size and pore throat size
distribution of the rock can be calculated from the mercury intrusion pressure and mercury
intrusion volume data. This will be discussed at further length in the section on capillary
pressure.

Dr. Paul Glover

Page 48

Formation Evaluation MSc Course Notes

5.3.4

Porosity

Gas Expansion

This method relies on the ideal gas law, or rather Boyles law. The rock is sealed in a
container of known volume V1 at atmospheric pressure P1 (Figure 5.5). This container is
attached by a valve to another container of known volume, V2, containing gas at a known
pressure, P2. When the valve that connects the two volumes is opened slowly so that the
system remains isothermal, the gas pressure in the two volume equalises to P3. The value of
the equilibrium pressure can be used to calculate the volume of grains in the rock Vs.. Boyles
Law states that the pressure times the volume for a system is constant. Thus we ca write the
PV for the system before the valve is opened (left hand side of Eq. (5.2)) and set it equal to the
PV for the equilibrated system (right hand side of Eq. (5.2)):

P1 ( V1 Vs ) + P2 V2 = P3 ( V1 + V2 Vs )

(5.2)

The grain volume can be calculated:

P V + P2 V2 P3 (V1 V2 )
Vs = 1 1
( P1 P2 )

(5.3)

In practice P1, P2 and P3 are


measured, with V1 and V2
known
in
advance
by
calibrating the system with
metal pellets of known volume.
The bulk volume of the rock is
determined
before
the
experiment by using either
vernier callipers and assuming
that the sample is perfectly
cylindrical,
or
after
the
experiment and subsequent
saturation
by
Archimedes
Method (discussed later), or by
fluid displacement using the
saturated sample. The bulk
volume and grain volume can
then be used to calculate the
connected porosity of the rock.
Any gas can be used, but the
commonest is helium. The
small size of the helium
molecule means that it can
penetrate even the smallest
pores.
Consequently
this

Dr. Paul Glover

Page 49

Formation Evaluation MSc Course Notes

Porosity

method gives higher porosities than either the imbibition or mercury injection methods. The
method itself is very accurate, insensitive to mineralogy, and leaves the sample available for
further petrophysical tests. It is also a rapid technique and can be used on irregularly shaped
samples. Inaccuracies can arise with samples with very low. Low permeability samples can
require long equilibration times in the helium porosimeter to allow diffusion of helium into
the narrow pore structures. Failure to allow adequate time will result in excessively high grain
volumes and low porosities.

5.3.5

Density Methods

If the rock is monomineralic, and the density of the mineral it is composed of is known, then
the pore volume and porosity can be calculated directly from the mineral density and the dry
weight of the sample. This method gives the total porosity of the rock, but is of no practical
use in petrophysics.

5.3.6

Petrographic Methods

This method is used to calculate the two dimensional porosity of a sample, either by point
counting under an optical microscope or SEM, or image analysis of the images produced from
these microscopes. Commonly a high contrast medium is injected into the pores to improve
the contract between pores and solid grains. This method can provide the total porosity, but is
wildly inaccurate in all rocks except those that have an extremely isotropic pore structure.
However, it has the advantage that pore types and the microtextural properties of the rock can
be determined during the process.

5.3.7

Other Techniques

Other techniques include porosity by (i) analysing all evolved fluids (gas+water+oil) and
assuming that their volume is equal to that of the pore space, (ii) CT scanning, and (iii) NMR
techniques.

5.3.8

Bulk Volume Measurement

Most of the methods reviewed above require the knowledge of the bulk volume of the rock
sample. Three ways are commonly used. These are (i) by using callipers, (ii) using fluid
displacement, and (iii) using Archimedes method.
Vernier Callipers If the rock is a perfect right cylinder with smooth surfaces, then calliper
measurements of length and diameter can give quite an accurate bulk volume. In this case
several measurements (approx. 10) are made of the length and the diameter, and the arithmetic
mean of each is used. Repeatability and accuracy then depend mainly upon surface texture of
the sample. Repeat helium expansion determinations of porosity on samples with smooth
surface textures where the calliper bulk volume is used should fall within 0.3 porosity
percent regardless of actual porosity. Inaccuracies can arise with samples with very high
permeability. High permeability sandstone samples are frequently friable, have large grain and
pore sizes, and do not produce smooth surfaced right cylinders when plugs are drilled and
accurate bulk volume determination becomes difficult. Straightforward measurement with

Dr. Paul Glover

Page 50

Formation Evaluation MSc Course Notes

Porosity

vernier callipers is not possible and Archimedes method or other liquid displacement methods
have to be used.
Fluid Displacement This method notes the displacement of fluid on a graduated scale when
the rock sample is placed in a container containing the fluid. If the fluid automatically enters
the pores errors will result. The method is commonly carried out with a non-wetting fluid such
as mercury, or with other fluids with a sample that has already been saturated. Mercury
displacement is carried out in a pyknometer fitted with a calibrated pump (Kobe method Figure 5.4). The sample is immersed in mercury during this test and will give erroneous
results where mercury enters samples with very large pores. There is also a tendency to give
high bulk volumes if air is trapped where the sample touches the top of the chamber. The
Kobe method is used as the first part of the mercury injection method (Section 5.3.3).
Archimedes Method The sample is weighed dry, fully saturated with formation brine whose
density is accurately known. The saturated sample is then weighed suspended under a balance
in air, and again while suspended in the fluid in which it is saturated. The various weight
readings, and the density of the fluid allow the bulk volume of any irregular sample to be
found accurately. The difference in the weight between the saturated sample suspended in air
and that when suspended in the fluid is equal to Vsf, where f is the density of the fluid.
There are few sources of significant error in this method, provided no fluid drains from the
plug whilst it is weighed in air. The most difficult part is judging how much excess fluid to
remove from the surface of the plug. Vuggy limestones present particular problems which
may only be overcome by whole core measurements. The contents of vugs exposed on the
plug surface may have been disturbed during drilling. Internal vugs may be partially filled
with solids from the drilling fluid during the coring process. If exposed vugs are genuinely
part of the pore volume, then bulk volume must be obtained by calipering since these will not
be taken account of by liquid immersion techniques.

Dr. Paul Glover

Page 51

Formation Evaluation MSc Course Notes

Porosity

It should be noted that these are all laboratory methods. Well logging utilises several other
different techniques, which all have larger errors associated with them. These are based on
acoustic, electromagnetic, NMR, and radioactive processes.

5.4 Influence of Stress


SCAL porosity measurements have to be done at overburden pressure if they are to be
correlated with downhole measurements. These measurements are made using the overburden
cell (Figure 5.6) attached to a helium expansion porosimeter. Pore volume changes can also
be observed whilst measuring formation factor at overburden pressures. It is not possible to
repeat determinations without allowing time for stresses in the core to be relieved. It is
conceivable that permanent damage could result when applying overburden to poorly
cemented cores, thus if a plug has to be used for a number of tests, overburden measurements
should form the later stages of the test sequence. As with routine poroperms, whole core
measurements may be necessary if samples are vuggy, fractured or contain stylolytes. The
precision of the data obtained is similar to that of routine poroperms. Care has to be taken
that samples are given sufficient time to allow compaction to occur at each overburden
pressure. The resulting porosity data is usually displayed as a fraction of that at ambient
pressure as function of overburden pressure (Figure 5.7), or as pore volume compressibility
(pore volume/pore volume/psi), as shown in Figure 5.8.

Dr. Paul Glover

Page 52

Formation Evaluation MSc Course Notes

Porosity

Dr. Paul Glover

Page 53

Formation Evaluation MSc Course Notes

Single Phase Permeability

Chapter 6: Single Phase Permeability


6.1 Introduction and Definition
Permeability is a property of a porous medium that characterises the ease which fluids flow
though it in response to an applied fluid pressure gradient.
The primary objective for permeability measurements applied to the hydrocarbon industry is
that they should be fit for purpose. In this case the purpose is to provide data that can be used
in accurate and effective reservoir modelling. If the reservoir model is to be used to help the
understanding of a dry gas reservoir at ambient conditions, then horizontal air permeability
measurements at ambient conditions will be fine. However, the reservoirs that are of interest
are rarely so simple, and it should be our aim to build multi-phase models capable of
modelling oil reservoirs at in situ conditions. Fluid permeabilities measured at or corrected to
relevant reservoir conditions using relevant fluids are essential inputs if such models are to be
representative of the reservoir. The fluid saturation and number of mobile fluids have a great
effect on permeability, reducing it below that for a dry rock containing a single fluid. This
section will deal with single phase fluid permeabilities. In particular the gas and Klinkenberg
permeability measurements that are made as part of RCAL, and the single phase liquid
permeabilities that are part of the more complex relative permeability SCAL tests, but which
are sometimes carried out on their own as part of RCAL.

6.1.1 Basic Definitions


Feynmann once said that, for a scientific measurement to be successful, the scientist or
engineer must know exactly what he or she is measuring. This comment has many
implications for the scientist. For the reservoir engineer/petrophysicist it requires that the
meaning of permeability is understood. So back to basics; permeability characterises the ease
with which fluids flows through a medium in response to a fluid pressure gradient. However,
permeability is not measured directly, but calculated from other physical measurements with
various theoretical and empirical relationships. The dependence on these relationships has the
implication that the resulting permeabilities are dependent on various assumptions and
boundary conditions. The relationship used in the hydrocarbon industry is the empirically
derived Darcys Law in 1856, derived using the apparatus shown in Figure 6.1.

q=KA

(P Pout )
h
= K A in
L
L

(6.1)

where: q = water flow rate


A = cross-sectional area of sand pack
L = length of sand pack
h = difference between the water heights in the manometers in Figure 6.1 (h1-h2)
K = A constant of proportionality characteristic of the sand pack (permeability)
Pin-Pout = fluid pressure gradient.

Dr. Paul Glover

Page 54

Formation Evaluation MSc Course Notes

Single Phase Permeability

The units of permeability used in the oil industry are the darcy, D, and the millidarcy, mD.
It is worth noting that the S.I. unit of permeability is in per metres squared (m-2), and shows
that there is an implicit spatial scaling of permeability in the measurement itself. This fact is
often overlooked when we use core measurements made at core plug scale (core volume
approximately 40 cm3), and then happily (and naively) compare it directly with logging
measurements, whose scale volume (volume of sensitivity) is 15000 cm3, and model reservoir
wide processes, whose scale volume may be approximately 1015 cm3! A permeability of 1 D
allows the flow of 1 cm3 per second of water with 1 centipoise, cP, viscosity, through a crosssectional area of 1 cm2, when a pressure gradient of 1 atmosphere pressure per centimetre is
applied. (1 D ~ 10-12 m-2.)
It should be understood that Darcys law, Eq. (6.1), was derived for unconsolidated sand
packs, assumes unreactive aqueous fluids with constant properties, and requires correction for
the different viscosity of different fluids, and correction for gas slippage (Klinkenberg effect)

Dr. Paul Glover

Page 55

Formation Evaluation MSc Course Notes

Single Phase Permeability

and inertial effects (Forchheimer effect) if used with gaseous fluids. In practice it is applied to
all rocks even though it is not clear that this is a valid extrapolation. One should, therefore,
always question the accuracy of a core permeability measurement. The law, Eq. (6.1), has
been extended for practical use in the following ways:

Inclusion of a fluid dynamic viscosity so that unreactive fluids other than brines can be
used.
Rewriting the h term in terms of absolute pressures.
Writing the flow rate, q, as volume flow per time (q=V/t).
Inclusion of a constant to take account of the units commonly used in measurement.

Thus the working equation for measuring single phase liquid permeabilities in the
hydrocarbon industry is:

K( mD) = 1000

L V
1
A t P P
1
2

(6.2)

If gas is used we must take account of the compressibility of the gas giving the working
equation for measuring single phase gas permeability in hydrocarbon industry RCAL:

K( mD) = 2000
where:
K

L
A
V
t
q
P1
P2
Patm

=
=
=
=
=
=
=
=
=
=

Patm
L V
A t P2 P2
2
1

(6.3)

permeability, (millidarcies, mD)


viscosity, (centipoise, cP)
plug length (cm)
plug cross section (cm2)
volume of fluid passed in t seconds (cm3)
time (seconds)
Flow rate, q=V/t, cm3s-1
inlet pressure (atmospheres absolute)
outlet pressure (atmospheres absolute)
atmospheric pressure (atmospheres absolute)

Gas permeability measurements are the most common RCAL permeability measurements.
These measurements suffer from two problems that are not encountered with liquid
permeabilities. These are the Klinkenberg and Forchheimer effects.

6.1.2 The Klinkenberg Effect


Darcys modified law for gases Eq. (6.3) is not applicable at low gas pressures (gas densities).
This is because, at such low pressures, the mean free path of the gas molecules become larger
than the pore dimensions. When this happens, the gas cannot be considered to be a continuous
medium and fluid mechanics cannot be used reliably. In practice the effect causes measured
permeabilities to be overestimated at low pressures (Figure 6.2a). In the low density (gas
pressure) limit the permeability is expressed as:

Dr. Paul Glover

Page 56

Formation Evaluation MSc Course Notes

K app = K L 1 +

Single Phase Permeability

(6.4)

Here the apparent or measured permeability Kapp is dependent on the so-called Klinkenberg
permeability KL, the gas pressure P and a constant known as the slip factor, . The standard
solution to the problem involves the following steps:

Repeating the measurement of gas permeability, Kapp, at four or five different gas inlet
pressures, P1 and gas outlet pressures, P2.
Calculating the mean gas pressure in the core for each determination; Pmean = (P1 + P2)/2.
Plotting Kapp against 1/Pm.

Dr. Paul Glover

Page 57

Formation Evaluation MSc Course Notes

Single Phase Permeability

The resulting plot is a straight line with a positive gradient (Figure 6.2b). The intersection of
the curve with the x-axis at 1/Pm = 0 gives KL . The Klinkenberg permeability is independent
of gas pressure, and is effectively the permeability of the gas as P, i.e. the permeability for
a near perfect liquid (an infinitely compressed near perfect gas). The values of apparent
permeability depend on the type of gas used even though their different viscosities are taken
into account in the calculation of apparent permeability. However, the Klinkenberg
permeability is independent on the type of gas used as all gases have the same properties in
the P limit (Figure 6.2c). This makes the Klinkenberg permeability very useful, for it can
be compared for different samples that had their gas permeabilities measured with different
gases at different gas pressures. The Klinkenberg permeability should be approximately the
same as the permeability of the rock when 100% saturated with a single phase reservoir liquid
such as water or oil. The gradient of the Klinkenberg plot gives the slip factor, which can be
used to characterise the rock microstructure.
The Klinkenberg correction should be applied to all core analysis measurements without fail.

Dr. Paul Glover

Page 58

Formation Evaluation MSc Course Notes

Single Phase Permeability

6.1.3 The Forchheimer Effect


At high gas flow rates (high differential pressures P1-P2), the gas accelerates through pore
throats and decelerates in pore bodies sufficiently for the gas inertia to cause turbulence.
Darcys law is an approximation of Navier-Stokes law, both of which require flow to be
laminar. Thus when the flow rate is fast enough for the flow to be turbulent, neither work. In
practice such high flow rates are avoided in all core analysis measurements. If they are
encountered they show up as underestimates in gas permeability measurements that are
recognised as an increase in the gradient of the K versus 1/Pm curve at low values of 1/Pm.

6.1.4 Averaging Permeabilities


It has been shown that the most probable permeability behaviour of a heterogeneous porous
medium made up of n randomly distributed regions of differing uniform permeabilities, K1 to
Kn, is described by the geometric mean of the individual permeabilities, which corresponds to
the mode of a log-normal distribution:

K g = n K1 K 2 K 3 K n

(6.5)

The analysis is extremely complex. However, it is possible to analyse two simple systems of
different permeabilities that occur within core analysis and reservoir systems. These are (i)
flow through linear beds in series, and (ii) flow through linear beds in parallel.
Linear Beds in Series. The system is shown in Figure 6.3a. The beds have a cross-sectional
area A that is constant. Each bed has a thickness, Ti , and a uniform permeability Ki. The
pressures at the contact between each of the beds, Pi. can be analysed thus:

(P1 P4 ) = (P1 P2 ) + (P2 P3 ) + (P3 P4 )

(6.6)

Now using Eq. (6.1) with h replaced by the pressure difference, and noting that the thickness
of the total unit T is equal to the sum of the individual beds T1 etc., we get:

q T3
q T1
q T2
qT
=
+
+
K A K1 A K 2 A K 3 A

(6.7)

Rearranging we find that the mean permeability is the harmonic average of the individual
permeabilities:

Kh = n

{Ti K i }

(6.8)

i =1

For example analysing Figure 6.3b, where three layers of equal thickness T=1 m have
permeabilities 1000 mD, 200 mD, and 1 mD, we get the mean permeability equals 2.98 mD!

Dr. Paul Glover

Page 59

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Single Phase Permeability

Page 60

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Single Phase Permeability

Page 61

Formation Evaluation MSc Course Notes

Single Phase Permeability

Clearly the permeability is controlled by the smallest permeability because all the fluids that
pass easily through the higher permeability layers are held up by the low permeability layer.
Linear Beds in Parallel. The system is shown in Figure 6.3c. The beds have a thickness T
that is constant. Each bed has a cross-sectional area to flow, Ai , and a uniform permeability,
Ki. The pressures at the inlet P1 and outlet P2 of the complete unit will be the same for all
layers, but each layer will transport a different fraction qi of the total flow rate qt thus:

q t = q1 + q 2 + q 3

(6.9)

Now using Eq. (6.1), with h replaced by the pressure difference and noting that the total area
A = A1+A2+A3, we get:

K A ( Pin Pout ) K1 A1 ( Pin Pout )


=
+
T
T
K 2 A 2 ( Pin Pout ) K 3 A 3 ( Pin Pout )
+
T
T

(6.10)

Rearranging we find that the mean permeability is the arithmetic average of the individual
permeabilities:

Ki Ai
K a = i=1

(6.11)

For example, analysing Figure 6.3d, where three layers of equal area A = 1 m2 have
permeabilities 1000 mD, 200 mD, and 1 mD, we get the mean permeability equals 400 mD!
The mean permeability falls much more into the mid range because the fluids partition for
flow into each of the layers depending on its permeability. In this case, the layer with the
highest permeability conducts 83.3% of the flow.
For comparison, the geometric mean of equal volumes of 1000 mD, 200 mD, and 1 mD is
10.6 mD, which falls between the two extreme cases analysed above, and represents random
arrangement of equal volumes of material with these three permeabilities.

6.1.5 Notes on RCAL Permeabilities


Gas permeabilities corrected for the Klinkenberg effect are commonly used, however this
measurement provides the most optimistic values of permeability mainly due to the
measurement being done for; (i) single phase gas fluids that are not representative of the true
reservoir fluids, (ii) low overburden pressures and temperatures that are not representative of
the in situ reservoir conditions, and (iii) cleaned dry rocks. Other measurement methods
account for these problems, but are more expensive, and often we are asked to use
Klinkenberg permeabilities where better measurements are unavailable. It is therefore
important for us to understand the factors affecting the determination of permeability

Dr. Paul Glover

Page 62

Formation Evaluation MSc Course Notes

Single Phase Permeability

measurements, such that the quality and relevance to the problem of any permeability dataset
can be assessed.
The factors affecting core permeability measurements fall into three broad categories; (i)
planning errors, (ii) sample errors, (iii) measurement errors, and (iv) analysis errors. Planning
errors are the fault of the person who commissions the permeability study. It is very tempting
to order a standard routine core analysis study. However, resources and time can be saved by
the commissioning manager thinking carefully about the purpose that the data is required for.
Klinkenberg permeabilities should not be used to estimate the efficiency of a waterflood, yet
some companies do so by correcting them to effective relative permeabilities using rules of
thumb that do not take account of the fluids and reservoir wettability adequately. Sample
errors are associated with; (i) sampling frequency, location, orientation, type and size; all of
which affect how representative the 40 cm3 sample is of the properties of the 1015 cm3 sized
reservoir, (ii) the type of drilling fluids, and (iii) the state of preservation and the process of
cleaning and drying, which can affect permeability greatly in shaly sandstones. Measurement
problems are related to the accurate measurement of pressure and flow, and are dependent
both on the initial experimental rig design as well as the permeameter operator. Finally,
Analysis problems involve the relevant use of the derived data and close the circle to the
planning stage. The indiscriminate lumping together of permeability data from different
measurement techniques, bad poroperm cross-plot analysis, and inefficient core-log
correlation of poroperm data, all contribute to inaccurate analysis, and almost always are the
result of either ignorance of the meaning and limitations of permeability data, or an effort to
make do with irrelevant data resulting from poor permeability study planning.

6.2 Controls on Permeability


6.2.1

Porosity

There have been several attempts to derive a general relationship between porosity and
permeability. In many ways, however, all attempts are bound to fail at a fundamental level
since porosity is a scalar measurement and permeability is a vector measurement. Clearly
though it is reasonable to assume that permeability should increase with porosity in
unfractured reservoirs without significant diagenetic. One of the most well known models
linking porosity and permeability is known as the Kozeny-Carman model that considers the
porous media to be made up of bundles of capillary tubes. The basic equation is:

K KC =
where:
KKC
c
d

=
=
=
=

c d2 3

(1 )

(6.12)

Kozeny-Carman predicted permeability, mD


A constant
Median grain size diameter, microns
Effective porosity

Despite the obvious invalid capillary tube assumptions, this model remains one of the best
predictors of permeability, and is often used in the hydrocarbon industry.
Dr. Paul Glover

Page 63

Formation Evaluation MSc Course Notes

Single Phase Permeability

Another commonly used empirical model is that of Berg:

K B = d 2 5.1

(6.13)

where KB is the predicted permeability. Although this empirical model has been concocted
from a range of rocks and it is clear that the equation may not work on samples from other
locations.
Recently, a new model has been proposed by Revil, Glover, Pezard and Zamora (RGPZ). This
is a non-empirical model that is derived from the fundamental understanding of the electrokinetic properties of rocks, and hold the potential for improved permeability prediction for
rocks of different porosities, grain sizes, and pore tortuosities. It is expressed as:

K RGPZ =
where:
KRGPZ
m
d
a

=
=
=
=
=

Dr. Paul Glover

d 2 3m
4am

(6.14)

RGPZ predicted permeability, D


The cementation exponent
Median grain size diameter, microns
Grain packing index
Effective porosity

Page 64

Formation Evaluation MSc Course Notes

Single Phase Permeability

Note that all of these models use a grain size diameter to scale the predicted permeability to
the size of rock microstructure, and to ensure that the models are dimensionally correct.
Figure 6.4 compares the three models described above for a range of clean sandstone, shaly
sandstone, and carbonate samples. Lines are placed at 1 mD in Figure 6.4; rocks with
permeabilities less than this value are considered to be non-reservoir rock (i.e. unproducible
economically).

6.2.2

Bedding

Permeability is a vector property, and as such, is greatly affected by directional heterogeneity


within core samples. The commonest cause of such heterogeneities is bedding. It is a general
rule that the vertical permeability within a reservoir (i.e. that perpendicular to the bedding) is
lower than that in the bedding plane (horizontal permeability). In fact the vertical permeability
is often about a third of that in the horizontal direction. It should also be noted that some of
the difference between the vertical and horizontal permeabilities results from differences in
the way the local stress fields in the vertical and horizontal directions compact pores and close
microcracks.

6.2.3

Pore Geometry

Permeability is highly dependent on the tortuosity of the pore fluid flow paths. Tortuosity can
be affected by many rock characteristics, including:

Grain size and its distribution


Grain shape
Sorting
Grain orientation
Packing arrangement
Degree and type of cementation
Amount, orientation and connectivity of micro-fractures
Clay content
Bedding
Diagenesis

The detailed relationships are known only qualitatively, and the relative importance of each
vary from rock type to rock type. For example, the permeability of carbonates is primarily
controlled by; (i) dissolution porosity, (ii) dolomitization, and (iii) fractures.

6.2.4

The Stress Conditions

Permeability is very sensitive to stresses that compact the rock. This compaction can occur in
any direction not just vertically. However, vertical compaction is usually the most important.
Indeed the local stress state may be such that dilatancy occurs (formation of fractures)
increasing the permeability of the rock. In all cases it is poorly consolidated rocks that are
affected to the greatest extent. Figure 6.5a and b show the effect of increasing the hydrostatic
confining pressure on the permeability of a rock. Figure 6.5c compares the effect of
overburden stress on permeability compared to the effect upon porosity. It can be seen that
overburden stress affects permeability much more than porosity. This is because permeability

Dr. Paul Glover

Page 65

Formation Evaluation MSc Course Notes

Single Phase Permeability

is very sensitive to the tortuosity of fluid flow paths through the rock, and such changes are
associated with very small changes to the rock porosity Overburden stress compacts the rock
pressing the grains together. The size of the pores reduces little, but the pore throats that
control the passage of gas between the pores undergo much greater closure, effecting the
permeability to a greater extent.
The large decreases
observed indicate that it
is very important to
apply corrections to
permeabilities measured
at
low
confining
pressures before they are
considered
to
be
representative of the
reservoir, or measure the
permeability at reservoir
stress conditions in the
first place (SCAL). It
should also be noted that
fracturing
(both
macroscopic and microfractures), that increases
the permeability of the
rock samples when
measured
in
the
laboratory,
can
be
caused by drilling and
concomitant upon the
sudden reduction in
stress experienced by the
rock upon extraction of
the core from the well.
These fractures can be
closed
again
by
measuring the rock at
reservoir conditions, but
it is very difficult to
know how to correct low
pressure
Klinkenberg
permeability
determinations for such
fracturing.

Dr. Paul Glover

Page 66

Formation Evaluation MSc Course Notes

Single Phase Permeability

6.3 Laboratory Determinations


6.3.1

Steady State Gas Permeability Determinations

Routine permeability measurements are made by confining plugs in Hassler core holders,
Figure 6.6, applying nitrogen pressure to one end and measuring flow rate and pressure
differential. Figure 6.7 shows a steady state gas permeability rig that is equipped to measure a
large range of permeabilities (i.e. gas flow rates). Standard hydrocarbon industry rigs look
similar, but have fewer options for measuring upstream pressure and flow rate.

Dr. Paul Glover

Page 67

Formation Evaluation MSc Course Notes

Single Phase Permeability

For plugs having moderate permeabilities, 5-500 mD, repeat determinations at given
confining, inlet and outlet pressures should fall within a few percent. Normally four or five
consecutive measurements are made at various mean pressures (Pm) to enable a Klinkenberg
plot (Figure 6.2b) of permeability vs. 1/Pm to be made. Extrapolation to infinite mean
pressure gives the equivalent liquid permeability, KL. Permeabilities above about 500 mD
become less precise as the measured pressure differential falls leading to higher experimental
errors. High permeabilities also imply large pores, large grains and rough surface texture.

Dr. Paul Glover

Page 68

Formation Evaluation MSc Course Notes

Single Phase Permeability

Very rough surfaces may need wrapping in soft PTFE tape or repair with epoxy to ensure
proper sealing by the Hassler sleeve. The sleeve pressure used will depend upon plug surface
texture and the hardness of the rubber sleeve. Low permeabilities (less than 5 mD) do not
normally present any problems; but for normal reservoir applications, a cut off value of 0.01
mD is applied. Values below this are simply reported as less than 0.01 mD, and are not
interesting as a reservoir. In practice rocks with permeabilities less than 1 mD are considered
to be non-reservoir rock (i.e. unproducible economically). If cap rocks are being investigated,
the actual permeability values will be reported, whatever their permeability. Caution is needed
in the handling of friable, poorly cemented samples. Gradual compaction can occur even with
sleeve pressures as low as 400 psi. Consequently long equilibration times may be necessary
for this type of sample. The first indication of this type of behaviour occurs when carrying out
the normal repeat timings of gas flow, when steadily decreasing flow rates are observed.

6.3.2

Unsteady State Gas Permeability Determinations

This is not as standard as the steady state method. It applied a volume of gas at a high initial
pressure to one end of the sample and then measures the decay of the pressure as the gas leaks
away through the core. One advantage of this method is that it can be used to determine the
permeability of very low permeability rocks. It has therefore been used to measure the
permeability of cap rocks. It must be said, however, that leakage through cap rocks is now
recognised to depend primarily on fractures through the cap rock rather than the permeability
of the bulk rock itself, and so these measurements are being done less and less.

6.3.3

Steady State Liquid Permeability Determinations

Permeabilities to oil and water at 100% saturation of each fluid, or of oil in the presence of Swi
can also be easily carried out. The saturated samples are placed in a core holder. The required
fluid is flowed through the sample, while measuring the steady state volume flow and pressure
differential (see Figure 6.8 for a schematic diagram of a typical permeameter set-up). All
fluids used should be degassed prior to use. The permeability is calculated from Eq. (6.2).
There is no need to institute a Klinkenberg correction, but the data is carefully examined to
ensure that the flow is laminar by carrying the test out at various flow rates and checking
whether they all give approximately the same permeability. Those high flow rates that are
suspected to contain turbulent (Forchheimer) effects are discarded.

Dr. Paul Glover

Page 69

Formation Evaluation MSc Course Notes

Single Phase Permeability

The values of permeability Kw at Sw=1 or Ko at So=1 should be approximately the same as the
Klinkenberg permeability, KL. Ko at Sw=Swi and So=1-Sw will be less that that at So=1.

6.4 Data Handling


The correlation of core and logging data enable reservoirs to be assessed for production
potential. The full description of this process is outwith the scope of this course. However, we
will briefly examine some of the issues related to the correlation of core measurements with
logs, and the use of permeability measurements in poroperm cross-plots.

6.4.1

Core-Log Comparison

The comparison of porosity and permeability data from core measurements and log methods is
important to ensure that there is good agreement between them, allowing the measurements to
be used in reservoir modelling with confidence. The process should compare the log and core
data on the same log-type display (Figure 6.9). It is usually clear whether one of the curves
needs to be depth shifted relative to the other. If a shift is necessary it is usually implemented
for the core data, as uncertainties in core depth occur when there is not 100% core recovery

Dr. Paul Glover

Page 70

Formation Evaluation MSc Course Notes

Single Phase Permeability

form the hole. Corrections of 10 to 20 m are not uncommon. When comparing the depth
corrected core and log data it is usually clear whether there is a good match between the two.
The degree of match is an average determination made by eye as the two measurements will
rarely be in very close agreement. This is because the measurements are made by widely
varying techniques, with varying scales of measurement. For example, a standard core plug
will have a volume of investigation of about 40 cm3, compared to approximately 15000 cm3
for a wireline tool. Additionally, the various methods measure different properties. For
example core porosities are usually measurements of effective porosities (with non-connected
porosity and clay bound water excluded, and usually avoiding fractures), whereas log derived
porosities are generally measurements of total porosity. This results in the log porosities being
generally a little higher than the log-derived porosities.

6.4.2

Poroperm Cross-plots

Permeability is of incredible interest to the hydrocarbon industry as it describes how profitable


fluids can be extracted from reservoirs. Clearly, any way of predicting permeability is of great
interest too. One of the fundamental processes that is applied to permeability data is to plot it
on a log-lin permeability-porosity diagram (Figure 6.10). Often there is a relationship within a
given rock unit, and differences between rock units can be useful in the analysis of the
reservoir. The main aims are:

Dr. Paul Glover

Page 71

Formation Evaluation MSc Course Notes

Single Phase Permeability

To estimate permeability where only porosity data is available (e.g. In Figure 6.10, a
conglomerate with a porosity of 13.3% has a permeability of 100 mD).
To establish a porosity cut-off below which the reservoir is unproductive (e.g. the
data in Figure 6.10 has porosity cut-offs of 5.3% in the conglomerate and 10.7% in the
sandstone, corresponding to a permeability of 1 mD..

This method is very powerful if used in an homogeneous formation, but can produce
remarkably erroneous results if carried out badly. There are a few points to bear in mind when
using cross-plots:

Some positive correlation between porosity and permeability exists for non-fractured, nonvuggy rocks with the same degree of diagenesis (Figure 6.11).
The estimation method is based on a mathematical correlation that only takes account of
porosity and permeability.
No other factors are taken account of (e.g. diagenesis, fractures, vugs).
The log permeability scale can generate large permeability errors.
The cross-plot should be done for each individual rock unit if the relationship is to be
valid. Doing a cross-plot for the whole reservoir is a waste of time. Individual cross-plots
for each mineralogy/lithology and/or based upon grain and pore size information from
mercury porosimetry. Individual permeability zones can also be delineated by plotting the
distribution of permeabilities on a lin-log plot. This results in a log-normal distribution for
well controlled permeability data (see section 6.1.4). If the distribution is unimodal
(Figure 6.12a) then a cross-plot for all the data will be valid. If the distribution is multimodal (Figure 6.12b) then a cross-plot must be done for the data belonging to each of the

Dr. Paul Glover

Page 72

Formation Evaluation MSc Course Notes

Single Phase Permeability

unimodal populations making up the multi-modal dataset (3 in the case of Figure 6.12b).
There are statistical tests that can distinguish which population a sample belongs to, e.g.
the Kolmogorov-Smirnoff test.
Some rocks do not produce a clear relationship (e.g. commonly carbonates).
There is no physical basis for this type of plot. In fact reference to Eqs. (6.12) to (6.13)
indicates that a log-log plot would be more appropriate, and Eq. (6.14) indicates that a linlin plot of permeability against 3m/m2 would provide the best results.

There are two other important points to bear in mind. First, for the cross-plot to be the most
valid, it should be done with permeabilities and porosities that are representative of the
reservoir. This means that the permeabilities should be Klinkenberg corrected, and both
permeabilities and porosities should be corrected to reservoir stress conditions. Any derived
permeabilities are then the permeabilities for complete saturation of the rock with the test
fluid, and will need to be reduced to relative permeabilities if required. If drill stem test
permeabilities are used, then these will have been made in the presence of multi-phase fluids,
and it is important to know more about the fractions of each fluid present and the wettability
of the rock before valid deductions can be made. The second point is that the porosities from
core measurements are effective porosities, whereas those from log measurements are
commonly total porosities. If data from both sets are to be used, then they must be reconciled
Dr. Paul Glover

Page 73

Formation Evaluation MSc Course Notes

Single Phase Permeability

before use. Figure 6.13 shows the typical ranges of poroperm relationships for various
lithologies and rock microstructure.

Dr. Paul Glover

Page 74

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Single Phase Permeability

Page 75

Formation Evaluation MSc Course Notes

Wettability

Chapter 7: Wettability
7.1 Introduction and Definition
There exists a surface tension between a fluid and a solid, in the same way that a surface
tension exists between an interface between two immiscible fluids (cf. the surface tension on
water under air that is sufficient to support the weight of a needle.). When two fluids are in
contact with a solid surface, the equilibrium configuration of the two fluid phases (say air and
water) depends on the relative values of the surface tension between each pair of the three
phases (Figure 7.1). Let us denote surface tension as , and solid, liquid and gas as s, l, and g
respectively. Each surface tension acts upon its respective interface, and define the angle at
which the liquid contacts the surface. This is known as the wetting (or dihedral) angle of the
liquid to the solid in the presence of the gas. Equilibrium considerations allow us to calculate
the wetting angle from the surface tensions:

lg cos = sg sl

(7.1)

This is known as Youngs equation (1805). Table 7.1 shows some contact angles and surface
tensions for common fluids in the hydrocarbon industry.
Table 7.1 Contact angles and interfacial tension for common fluid-fluid interfaces
Interface

Air-Water
Oil-Water
Air-Oil
Air-Mercury

Dr. Paul Glover

Contact Angle, ,
degrees

cos

Interfacial Tension,
dynes/cm

0
30
0
140

1.0
0.866
1.0
-0.765

72
48
24
480

Page 76

Formation Evaluation MSc Course Notes

Wettability

Note that:

If sg > sl , then cos > 0 and <90o.


If sg < sl , then cos < 0 and >90o.

For a stable contact | cos | 1, or equivalently | sg - sl | gl . This inequality is not


satisfied when lg + sl < sg, when liquid covers the whole solid surface. Alternatively, when
lg + sg < sl, the gas displaces the liquid away from the surface completely. Figure 7.2 shows
a range of different wetting conditions.

When one fluid preferentially covers the surface, it is called the wetting fluid, and the other
fluid is called the non-wetting fluid. The origin of these surface tensions arises in the different
strengths of molecular level interactions taking place between the pairs of fluids. For example
a quartz sandstone grain generally develops greater molecular forces between itself and water
than between itself and oils. Clean sandstones are therefore commonly water wet. However,
should that same grain have been baked at high temperatures in the presence of oil at depth,
then (i) its surface chemical structure can be altered, or (ii) the surface itself can become
coated, in such a way that results in the grain having greater molecular interaction with oils
than water, and hence become oil wet. IMPORTANT NOTE Properly measured wettabilities
on well preserved core and core plugs should form the initial step in choosing relative
permeability test methods.

Dr. Paul Glover

Page 77

Formation Evaluation MSc Course Notes

Wettability

7.2 Laboratory Determination


There are several methods for determining wettability of a rock to various fluids. The main
ones are:
(i) Microscopic observation This involves the direct observation and measurement of
wetting angles on small rock samples. Either a petrographic microscope or SEM fitted with an
environmental stage is used. The measurements are extremely difficult, and good data relies
more on luck than judgement.
(ii) Amott wettability measurements This is a macroscopic mean wettability of a rock to
given fluids. It involves the measurement of the amount of fluids spontaneously and forcibly
imbibed by a rock sample. It has no validity as an absolute measurement, but is industry
standard for comparing the wettability of various core plugs.
(iii) USBM (U.S. Bureau of Mines) method. This is a macroscopic mean wettability of a
rock to given fluids. It is similar to the Amott method but considers the work required to do a
forced fluid displacement. As with the Amott method, it has no validity as an absolute
measurement, but is industry standard for comparing the wettability of various core plugs.
(iv) NMR longitudinal relaxation and other wettability methods. These are briefly
summarised in Table 7.2.
Table 7.2 Summary of Wettability Measurement Techniques
Measurement Technique

Physical Observation

Amott and Amott-Harvey

Amounts of oil and water imbibed by a


sample spontaneously and by force.
Work required to imbibe oil and water.
Microscopic examination of the interaction
between the fluids and the rock matrix.
Changes in longitudinal relaxation time.
The distribution of grains at water/oil or
air/water interfaces.
Displacement of the non-wetting fluid from a
glass slide.
Shape and magnitudes of Kro and Krw curves.
Resistivity logs before and after injection of a
reverse wetting agent.
Adsorption of a dye in an aqueous solvent.

U.S. Bureau of Mines (USBM)


Microscopic examination
Nuclear Magnetic Resonance
Flotation method
Glass slide method
Relative permeability method
Reservoir logs
Dye adsorption

The first two methods are the commonest within the oil industry and are described in greater
detail in the next two sections.

Dr. Paul Glover

Page 78

Formation Evaluation MSc Course Notes

7.2.1

Wettability

Amott Wettability Measurements

Wettability measurements by the Amott method give a guide to the relative oil or brine
wetting tendencies of reservoir rocks. This can be crucial in the selection of relative
permeability test methods to generate data relevant to the reservoir situation. It is not always
possible to reproduce reservoir wettabilities in room condition relative permeability tests.
However, an appreciation of the difference between reservoir and laboratory wettabilities can
assist in interpretation of laboratory waterfloods.
The Amott method (Figure 7.3) involves four basic measurements. Figure 7.4 shows the data
produced with the water wetting index given by AB/AC and the oil wetting index by CD/CA.
(i)
(ii)
(iii)
(iv)

The amount of water or brine spontaneously imbibed, AB.


The amount of water or brine forcibly imbibed, BC.
The amount of oil spontaneously imbibed, CD
The amount of oil forcibly imbibed, DA

Figure 7.4 shows the initial conditions of the sample (point X) to be oil saturated at Swi. The
spontaneous measurements are carried out by placing the sample in a container containing a
known volume of the fluid to be imbibed such that it is completely submerged (steps 1 and 3
in Figure 7.3 for water and oil respectively), and measuring the volume of the fluid displaced
by the imbibing fluid (e.g. oil in step 1 of Figure 7.3). The forced measurements are carried
out by flowing the imbibing fluid through the rock sample and measuring the amount of the
displaced fluid (steps 2 and 4 in Figure 7.3), or by the use of a centrifuge. The important
measurements are the spontaneous imbibitions of oil and water, and the total (spontaneous
and forced) imbibitions of oil and water. Water-wet samples only spontaneously imbibe water,

Dr. Paul Glover

Page 79

Formation Evaluation MSc Course Notes

Wettability

oil-wet samples only spontaneously imbibe oil, and those that spontaneously imbibe neither
are called neutrally-wet. The wettability ratios for oil (AB/AC) or water (CD/CA) are the
ratios of the spontaneous imbibition to the total imbibition of the each fluid.

Dr. Paul Glover

Page 80

Formation Evaluation MSc Course Notes

Wettability

In general use the samples to be measured are centrifuged or flooded with brine, and then
flooding or centrifuging in oil to obtain Swi. The standard Amott method is then followed. At
the end of the experiment the so called Amott-Harvey wettability index is calculated:

SpontaneousWaterImbibition SpontaneousOilImbibition

TotalWaterImbibition
TotalOilImbibition
AB CD
=
AC CA

Index=

(7.2)

Wettability indices are usually quoted to the nearest 0.1 and are often further reduced to
weakly, moderately or strongly wetting. The closer to unity the stronger the tendency.

7.2.2

USBM Wettability Measurements

This method is very similar to the Amott method, but measures the work required to do the
imbibitions. It is usually done by centrifuge, and the wettability index W is calculated from
the areas under the capillary pressure curves A1 and A2:

W=log

A1
A2

(7.3)

where, A1 and A2 are defined in Figure 7.5. Note that in this case the initial conditions of the
rock are Sw=100%, and an initial flood down to Swi is required (shown as step 1 in Figure 7.5),
although either case may be necessary for either the Amott or USBM methods. Figure 7.6
shows typical USBM test curves for water wet, oil wet and neutrally wet cores.

7.2.3

USBM and Amott Technique Comparison

Warning. While both methods are common within the oil industry, they show remarkable
differences especially near the neutral wettability region. In general, the Amott method is
probably the most reliable and accurate especially in the neutral wettability region. A
comparison of the two methods, together with contact angles, is given in Table 7.3.
Table 7.3 Comparison of the Amott and USBM Wettability Methods

Amott wettability index water ratio


Amott wettability index oil ratio
Amott-Harvey wettability index
USBM wettability index
Minimum contact angle
Maximum contact angle

Dr. Paul Glover

Oil Wet

Neutral Wet

Water Wet

0
>0
-1.0 to -0.3
about -1
105o to 120o
180o

0
0
-0.3 to 0.3
about 0
60o to 75o
105o to 120o

>0
0
0.3 to 1.0
about 1
0o
60o to 75o

Page 81

Formation Evaluation MSc Course Notes

7.3

Wettability

Sample Preservation and the Effect on Wettability

Sample preservation, preparation, storage and test conditions are important since wettability is
sensitive to oxidation of crude oil and to temperature. Preservation of samples at the well-site
is generally by wrapping in foil and wax coating. The process should be carried out as soon
as possible after removal of the core from the barrel. Samples are drilled and trimmed with
deaerated, simulated depolarised kerosene prior to testing. Tests carried out to assess
reservoir wettability must be made on preserved core at a temperature high enough to ensure
that any wax present in the residual core remains in solution. If the wax precipitates, it will
tend to increase the oil wetting tendency of the core. Room condition wettabilities may only
apply to reservoirs containing wax free crude or cleaned cores from laboratory tests.

Dr. Paul Glover

Page 82

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Wettability

Page 83

Formation Evaluation MSc Course Notes

Capillary Pressure

Chapter 8: Capillary Pressure


8.1 Introduction and Theory
Capillary pressure data are required for three main purposes:

The prediction of reservoir initial fluid saturations.


Cap-rock seal capacity (displacement pressures).
As ancillary data for assessment of relative permeability data.

Capillary pressures are generated where interfaces between two immiscible fluids exist in the
pores (capillaries) of the reservoir rock. It is usual to consider one phase as a wetting phase
and the other as a non-wetting phase. However, intermediate cases occur which can greatly
complicate the picture. The drainage case, i.e. a non-wetting phase displacing a wetting phase
applies to hydrocarbon migrating into a previously brine saturated rock. Imbibition data is the
opposite to drainage, i.e. the displacement of a non-wetting phase by a wetting phase. Thus,
the drainage data can usually be used to predict non-wetting fluid saturation at various points
in a reservoir, and the imbibition data can be useful in assessing the relative contributions of
capillary and viscous forces in dynamic systems.
The basic relationship (Figure 8.1) between capillary pressure, interfacial tension, contact
angle and pore radius is given by

Cp =

2 cos
.A
a

(8.1)

where;
Cp

a
A

=
=
=
=
=

capillary pressure (psi)


interfacial tension (dynes/cm)
contact angle (degrees)
pore radius (microns)
145 x 10-3 (constant to convert to psi)

Applying this to a water wet rock having a broad spectrum of pore entrance radii, oil
migrating into water filled pores under a given pressure differential will only enter pores
larger than those indicated by a in Equation (8.1). Thus for oil introduced at 2 psi into a
system having = 40 dynes/cm and = 0 (water wet), oil will only enter pores larger than

2 =

2 x 40 x 1
x 145 x 10 3
a

(8.2)

giving a = 5.8 m.
If capillary pressure data are available for a given system, it should be possible to convert to
another system of known and . This is expanded in Section 8.2.

Dr. Paul Glover

Page 84

Formation Evaluation MSc Course Notes

Capillary Pressure

8.2 Mercury Injection


These tests can only be carried out on cleaned, dried test plugs. They are initially immersed in
mercury at <10-3 mm Hg vacuum within the apparatus sketched in Figure 8.2. The pressure in
the system, effectively the differential across the mercury/vacuum interface, is raised in stages.
The volume of mercury which has entered the pores at each pressure is determined from
volumetric readings, and the proportion of the pore space filled can be calculated. This gives
the curve shown in Figure 8.3. Further readings can be taken as the pressure is lowered to
provide data for the imbibition case from Swi to Cp = 0 at point A in Figure 8.3.
The volume of mercury injected into the pores at a given pressure is usually expressed as a
proportion of the total pore space, and is presented as a pore size distribution (Figure 8.4) or
converted to oil or gas-brine data using appropriate contact angles and interfacial tensions.
Typical conversions are given below:-

C p (gas - brine) = C p (Air - Hg)

72 cos 0o
480 cos 130

(8.3)

= 0.233 C p (Air - Hg)

Dr. Paul Glover

Page 85

Formation Evaluation MSc Course Notes

C p (oil - brine) = C p (Air - Hg)

Capillary Pressure

25 cos 30o
480 cos 130

(8.4)

= 0.070 C p (Air - Hg)

Further conversion to height above oil, or gas-water contact is possible from

C p = hg ( w - o )

(8.5)

Displacement pressures for the assessment of cap rock seal capacity can be assessed from the
capillary pressure curves. Care has to be used and allowance made for the effect of surface
irregularities. This is especially true of samples with small total pore volumes, i.e. less than 1
ml that are typically the case with cap rock samples.
Approximate permeabilities can also be calculated from pore size distribution data; but care
has to be taken to exclude the contribution of surface irregularities, since a small number of
large pores disproportionately increases the calculated permeability. The data presented in
Figure 8.4 shows the effect of surface irregularities at pore throat diameters greater than 10
microns.

Dr. Paul Glover

Page 86

Formation Evaluation MSc Course Notes

Capillary Pressure

Mercury injection has the following disadvantages:


(i) It is a destructive test.
(ii) It is carried out on dried core which allows for no fluid-surface interactions.
(iii) It can cause collapse of accumulations of grain surface coating minerals.
The latter two effects lead to low implied connate water saturations. The data obtained apply
to a system containing a fully wetting phase and a fully non-wetting phase. The capillary
pressure data obtained will not necessarily apply to pores containing fluids showing partial
wetting preferences.
The advantages of the technique are that it is rapid (about three hours per sample) and
irregular samples can be used. A drainage capillary pressure curve can be produced in a
matter of hours and in certain circumstances drill cuttings can provide useful data.

Dr. Paul Glover

Page 87

Formation Evaluation MSc Course Notes

Capillary Pressure

8.3 Porous Plate Method


8.3.1

Gas-Brine Systems

This technique is generally applied in the drainage mode to air-brine systems starting with test
plugs which are initially brine saturated. The capillary pressure is applied across the test plug
and a brine saturated porous plate. The high displacement pressure of the porous plate allows
brine from the plug to pass through, but prevents flow of the displacing fluid (normally air).
The apparatus is shown in Figure 8.5. Plugs are removed at intervals and weighed until
weight (and therefore fluid) equilibrium is attained. The pressure applied is then increased
and the process repeated until a full curve of about six points is obtained.
In this method care has to be taken to maintain good capillary contact between the test plug
and the porous plate. This is assisted by using a paste of filter-aid and brine between the plate
and a filter paper. The test plug is positioned on the paper and a lead weight placed on the
plug to keep it solidly in place. There is also the danger that the water in the sample will be

Dr. Paul Glover

Page 88

Formation Evaluation MSc Course Notes

Capillary Pressure

evaporated by the gas pressure. To avoid this the input gas can be saturated with water by
bubbling it through a reservoir of water prior to use, and keeping a beaker of water inside the
porous plate pressure vessel.

The resulting data is presented as (i) air-brine capillary pressure versus brine saturation
(Figure 8.6), (ii) converted to oil-brine data (Figure 8.7), or (iii) as saturation versus height
above oil (or gas) - water contact.

Dr. Paul Glover

Page 89

Formation Evaluation MSc Course Notes

Capillary Pressure

The disadvantages of the porous plate method are that it is time consuming, needing up to
twenty days for equilibration at each pressure. Also, capillary contact with the porous plate
may be lost at high pressures. This causes erroneously high connate water saturations to be
implied. Imbibition measurements are not generally attempted.
The advantage of the method is that the test plug has at least one representative fluid in place,
i.e. the brine. This ensures that brine mineral interactions e.g. clay swelling, which affect pore
size and surface states, are taken account of. This is a large advantage over the mercury
method, which cannot take account of clay-water interactions.

8.3.2 Oil Brine Systems


It is more difficult to make oil-brine measurements than air-brine. For a fully water-wet
system, oil-brine drainage capillary pressure data can be inferred from air-brine data. Actual
oil-brine drainage capillary pressure can be measured using a Hassler cell fitted with a brine
saturated disc (see Figure 8.8). An initially brine saturated test plug is subjected to an oil
pressure at the inlet face, and the volume of brine produced (oil taken up) observed at the
outlet. Once equilibrium is achieved (this may take 1-3 weeks), the pressure is increased and
the process repeated until a full curve is obtained. Care has to be taken that the displacement
pressure of the disc is not exceeded, leading to brine displacement from the disc as well as the
test plug. This can only be checked and allowed for by weighing the disc before and during
the course of the experiments.
The disadvantages of the method are basically the same as for the gas-brine case, but with
added complications if test plug wetting characteristics differ from those of the reservoir. The
method is limited to approximately 50 to 100 psi depending upon disc characteristics.
Imbibition measurements are extremely difficult and are limited to about 25 psi.

Dr. Paul Glover

Page 90

Formation Evaluation MSc Course Notes

Capillary Pressure

The advantages of the method are that, if the wetting characteristics are as in the reservoir,
then representative data should be obtained for both mixed and oil wet systems.

8.3.3

Other Methods

There are several other more advanced methods of measuring capillary pressure that are
currently being investigated by the hydrocarbon industry. All use a core in a Hassler or similar
coreholder under confining pressure. The major ones are:
(i) The Dynamic Method This involves injecting the two fluids into a rock core
simultaneously, and producing one behind a semi-permeable membrane.
(ii) The Semi-Dynamic Method This involves injecting a single fluid, while a membrane
at the far end of the rock is washed with the other fluid. This method can be used to measure
the complete drainage and imbibition parts of the capillary pressure curve.
(iii) The Transient Method This method is technologically complex, and involves the
measurement of saturation and pore fluid pressures simultaneously during fluid injection into
the sample.

Dr. Paul Glover

Page 91

Formation Evaluation MSc Course Notes

Capillary Pressure

8.4 Centrifuge Method


The centrifuge method is widely advocated by US companies. It relies upon increasing the gterm in the equation;

Cp = h g ( w - o ) ,

(8.6)

by spinning the core plug at a known radius and rpm. The average capillary pressure is then
given by:

C p ( psi) = 7.9x10 8 ( 1 2 ) R 2 ( rb2 rt2 )

(8.7)

where 1 and 2 are the densities of the two phases present and rb and rt are the radii of
rotation of the bottom and top of the core respectively.
For the oil-brine drainage cycle, brine saturated test plugs are immersed in oil in specially
designed holders. Starting at a low rpm setting, the amount of brine expelled from the plug is
noted for a given rate of rotation. The volumes are measured in the following manner. A
calibrated glass vial is attached to the end of the sample. The volume of fluid being deposited
in this vial can be read while the centrifuge is spinning fast using a stroboscope. The rate is
then increased in stages and produced brine volumes are recorded for each rotation speed to
give the drainage curve. The imbibition curve can then be followed by stopping the centrifuge
allowing spontaneous imbibition to occur to point A at Cp = 0 (Figure 8.3). The fluid in the
imbibition cell is then changed from brine to oil, and the portion of the curve from A (Cp) = 0
to Sor (Figure 8.3) can then be followed by recording the volume of oil produced at several
increasing rates of rotation.
The main disadvantage of the centrifuge method is that a capillary pressure gradient is applied
which must inevitably give rise to a saturation gradient. This will be more exaggerated at low
pressures.
The advantages of the centrifuge method are that it is rapid, a full drainage and imbibition
cycle being complete in a matter of days, and that oil-brine data can be obtained, hopefully,
under representative wetting conditions. Centrifuges can also be operated at elevated
temperatures (up to 150).
Differences in wetting characteristics should be taken into account when applying laboratory
data to the field. Thus air-brine and mercury-air data obtained on cleaned core will represent a
fully water wet system for the drainage case. They may not adequately describe a mixed or oil
wet system. The closest one can get to this situation, on a routine basis, is with an actual oilbrine system when one would expect to find lower Swi values than a water wet system.

Dr. Paul Glover

Page 92

Formation Evaluation MSc Course Notes

Capillary Pressure

8.5 Applications of Methods


The applications of the various major methods described above are given in Table 8.1.
Table 8.1 Parameters associated with various capillary pressure measurement methods.
Range

Method
Mercury
Hg - Air

Drainage
Imbibition

Porous Plate
Air - Brine
Oil - Brine

Sw = 1 to Swi
Swi to Cp=0, Cp=0 to Sor
Cp=0 to Sor
As sample

Swi to Cp=0

Sample wetting
state
Sw precision
Pressure limit, psi
Elapsed time

Centrifuge
Air - Brine
Oil - Brine

Water wet
tends low
1150
1-2 days

tends high
70-100
3 months

tends high
70-100
3 months

not known
1000-2000
1-3 days

Cp=0 to Sor

not known
1-3 days

8.6 Capillary Pressure Prediction (Leverett Method)


The Leverett J function is an attempt to correlate capillary pressure with pore structure
(defined in terms of porosity and permeability). The basic capillary model predicts that to
displace a wetting phase with a non-wetting phase then:

Cp =

2 Cos
.
a

(8.8)

For a simple capillary model, the mean hydraulic radius, a, is given by:

K
a=2 2

12
(8.9)

Combining Eq. (8.8) and (8.9) we get;

12

K
Cp

cos

1
2

(8.10)

from which the Leverett J function is defined:

Dr. Paul Glover

Page 93

Formation Evaluation MSc Course Notes

Capillary Pressure

12

K
Cp

J=
cos

(8.11)

or;

K
J = 0.217 C p

12
(8.12)

where Cp is in psi, is in dynes/cm, and K is in mD.


Given a typical J curve, the capillary pressure curve for a material of similar pore structure can
be calculated for a given value of and K.

Dr. Paul Glover

Page 94

Formation Evaluation MSc Course Notes

Electrical Properties

Chapter 9: Electrical Properties


9.1 Introduction
Laboratory measurements of electrical properties (formation factor and resistivity index) are
intended to complement those made during down hole logging operations. The data are used
to refine values of n, a and m in Eq. (9.1) and (9.2) below:

F =

Ro
a
= m
Rw

(9.1)

where;
F
Ro
Rw
a

=
=
=
=
=
=

Formation factor
Resistivity of brine saturated rock
Resistivity of brine
Constant
Porosity
Cementation exponent

and,

I=

Rt
1
=
R o S nw

(9.2)

where;
I
Rt
n

=
=
=

Resistivity index
Resistivity of rock at Sw < 100%
Saturation exponent

Without prior laboratory data, n is generally assumed to have a value of 2. Typically,


laboratory derived data gives values between 1.7 and 2.4 (Figure 9.1).
Equations (9.1) and (9.2) are empirical in nature. They are generally adhered to by clean
samples, but where clays (usually described as shales in this context) are present large
deviations can occur. Empirical corrections for shale effects, i.e. for m to m* and n to n* are
possible but the best procedures are in doubt. The shale effect, which is primarily due to
enhanced surface conduction in the high surface area clays, can now also be corrected for
using fundamental theory, but the governing equations for the physical processes involved are
complex.
As with other SCAL tests, samples should be prepared and brought to the initial brine
saturated state without drying. The displacing phase used in resistivity index measurements is
air, thus these tests are carried out in conjunction with air-brine capillary pressure
measurement, most commonly using the porous plate method.

Dr. Paul Glover

Page 95

Formation Evaluation MSc Course Notes

Electrical Properties

9.2 Formation Factor


Formation factor is a function of porosity and pore geometry and is defined above in Eq. (9.1).
Laboratory measurements are made at room conditions or at overburden conditions, the
apparatus is sketched in Figures 9.2 and 9.3 respectively. Normally plugs require 2 weeks for
equilibration with formation brine. For room condition tests, measurements are made at
intervals of a few days until constant values are obtained. In the case of overburden
measurements, it is possible to flow brine through the plug in the cell and equilibrium is
achieved more rapidly.
Overburden resistance measurements are made at increasing pressures, but time has to be
allowed for equilibrium to be achieved. This is generally due to varying rates of compaction
under overburden pressures and, in the case of low permeability samples, it can take many
hours for brine to be fully expelled. Rates for compaction vary according to material, 24 hours
often being required for full equilibration. Repeat determinations cannot be made without
allowing plugs sufficient time to relax back to the unconfined state. This can take many weeks
and in some cases, where the pore structure is irreversibly damaged, the rock never returns to
its initial state. Once a rock has been used for resistance measurements at high overburden
pressures, it should never be used for further study for this reason. Overburden formation
factor measurements are generally combined with pore volume compressibility
determinations.

Dr. Paul Glover

Page 96

Formation Evaluation MSc Course Notes

Electrical Properties

The actual electrical measurements of resistance in themselves are very precise (better than
0.1%), but care has to be exercised in setting up the plug for the measurements. The most
important thing is that the electrodes at each end of the sample do not have any electrical
connection except through the rock. This may seem obvious and trivial to arrange, but this one
proviso causes the greatest difficulty when setting up a plug to be measured either at high
overburden pressures, during rare triaxial deformation experiments, or using high electrical
frequencies. The first and second of these situations arises from the need to electrically isolate
the electrodes at each end of the sample from each other and the pressure vessel. Insulation is
shown schematically in Figure 9.3, but in reality it is not always easy to arrange a robust
electrically insulating and pressure-proof leadthrough. In the last situation, it is the
capacitance of the leadthroughs and the pressure vessel itself that causes the problems. Even
though there is no direct conductive connection, high frequency current can leak from one
electrode to the other through the body of the pressure vessel by charge induction, even
though there is an insulator in between.

Dr. Paul Glover

Page 97

Formation Evaluation MSc Course Notes

Electrical Properties

High quality in-phase and out-of-phase resistance (conduction, and permittivity)


measurements can be made anywhere in the frequency range from DC to 10 MHz using
impedance analysers. The whole range of frequencies is generally used in academic rock
physics, where the technique is called impedance spectroscopy. In the oil industry, only one
frequency is used. This is usually 1 kHz, or near it, and is usually taken as the frequency at
which the out-of-phase component of the resistance is minimised, i.e. the frequency at which
the conduction is most ohmic. Note that this frequency avoids the highest frequencies where
current leakage can be a problem. It also avoids the low frequencies, where electrode
polarisation can be a problem.
The electrodes are commonly made of platinum gauze, upon which a fine dendritic structure
of amorphous black platinum has been electro-deposited. This increases the surface area of the
electrodes by several orders of magnitude, helping to reduce electrode polarisation effects to
negligible values. Often a pad of filter paper soaked in the pore fluid is inserted between the
rock and the platinum electrode to (i) homogenise the current flow, (ii) improve the electrical
connection, and (iii) avoid conductive minerals channelling current into the rock.
For room condition tests, care has to be taken to remove excess surface moisture from the
plug (so that the conduction along the plug surface is not measured) and to ensure that just
sufficient brine is contained in the electrode ends to give good contact. Contact problems are

Dr. Paul Glover

Page 98

Formation Evaluation MSc Course Notes

Electrical Properties

less pronounced in the overburden cell as brine is flowed through the system before
measurements are taken, and good electrical contact is assisted by the hydrostatic load on the
end pieces. Plug surface conduction is removed by placing the plug in a rubber sleeve, which
is squeezed tightly onto the plug by the confining pressure.
The main measurement is sample resistance, r. This is clearly dependent upon the length, L,
and the cross sectional area, A, of the sample. In order to compare samples, the resistance,
length and cross sectional area of the sample are used to calculate the resistance per unit
length and per unit cross sectional area of the sample rock; this is called the resistivity, R:

R = r

A
L

(9.3)

Note from Eq. (9.1) that the resistivity of the pore fluid is also required. This can be done in a
standard dip cell, but this method is prone to large systematic errors. More commonly a
specially designed fluid cell is used (Figure 9.4). This cell is connected to the same impedance
analyser as used for the main measurements, and at the same frequencies. The fluid resistance
obtained in this way is converted to a fluid resistivity by multiplication by a cell constant, that
varies from cell to cell, and with temperature and pressure. The cell constant is obtained by
calibrating the cell with fluids of accurately known composition and resistivity.

The precision of measurement is dependent upon operator skills and sample permeability. For
lower permeability samples (with small pores) repeat determinations should fall within a few
percent. For friable, high permeability samples repeatability is poorer, and in extreme cases
room condition tests may be impossible due to too rapid drainage of fluid from the sample.

Dr. Paul Glover

Page 99

Formation Evaluation MSc Course Notes

Electrical Properties

Cementation exponent, m, for a particular sample can be calculated directly from Eq. (9.1) if
it is assumed that a=1. More commonly, cementation exponent for a group of samples is
calculated graphically using a minimum of 10 samples covering a wide a permeability range
as possible. A typical data set is shown in Figure 9.5. Equation (9.1) can be rewritten as:

log F = log a - mlog

(9.4)

A log-log plot of formation factor against porosity gives a straight line, with a gradient equal
to the negation of the cementation exponent, and with a y-intercept at =1 equal to log (a). It
is common to see cementation data from a best fit through the data giving both m and a, but
most commonly the linear regression is forced through (=1, F=1) corresponding to a=1.

Some typical overburden measurements are shown in Figure 9.6.

Dr. Paul Glover

Page 100

Formation Evaluation MSc Course Notes

Electrical Properties

9.3 Resistivity Index


This involves similar measurements to formation factor except that resistance is measured at
values of Sw less than 100%. Plots of resistivity index and Sw give n, the saturation exponent
(Figure 9.7). The methods of desaturation include porous plate and centrifuge, thus resistivity
index measurements are conveniently combined with air/brine capillary pressure
measurements.
Note that Eq. (9.2) can be written as:

log I = 1 - n log Sw

(9.5)

So the saturation exponent, n, is the negation of the gradient of the log I versus log Sw plot
(Figure 9.7 shows typical data), and that the line should always pass through (Sw=1, I=1).
The practical considerations for these electrical measurements are similar to those for
formation factor, except that the tendency for brine to drain from the plug is essentially
removed because of the lower saturations. The electrical measurements can be performed to
better than 0.1% and saturation changes determined to 0.5 Sw%.
Care has to be taken to avoid evaporation losses during desaturation and measurement of
resistance. Plugs are stored in closed weighing bottles and only removed for the minimum
possible period.

Dr. Paul Glover

Page 101

Formation Evaluation MSc Course Notes

Electrical Properties

9.4 Shale Effects


Very little data is available on the reproducibility of m and n. Repeat determinations of n by a
skilled operator should result in good agreement, probably within a few percent.
Plots of Sw versus resistivity index often give very good straight lines, but this is not always
the case. Some scatter is likely at high values of Sw (above 80%) and at low values (below
20%).
The reasons for this are not well understood. Scatter at high values of Sw is difficult to
account for; but at low values of Sw it could be due to genuine clay effects, i.e. excess
conductivity of clays, or to water vapour loss and concentration of brine in the test plug.
Correction of m and n to m* and n* is still under general discussion within the industry.
These are generally corrected from cation exchange capacity (CEC), which can be obtained
from laboratory measurements on crushed rock. Unfortunately, the values obtained for CEC
can depend upon the amount of crushing the core has been subjected to, and the method used.
It is possible that measurement of formation factor at various salinities can give a better guide
to correction of m and n. A plot of core conductivity against brine conductivity (Figure 9.8)
can give a value for BQV/F*, which has then at least been obtained from direct electrical
measurements on intact core with the fluids of interest in place.

Dr. Paul Glover

Page 102

Formation Evaluation MSc Course Notes

Electrical Properties

This topic requires deeper discussion than can be covered here, but it should be borne in mind
that the relationships between electrical properties and saturations are empirical; linear
relationships cannot always be expected.
A more comprehensive discussion on the accuracy of electrical measurements are given in

references [2] and [3].

Dr. Paul Glover

Page 103

Formation Evaluation MSc Course Notes

Relative Permeability

Chapter 10: Relative Permeability


10.1

Introduction and Definition

Routine permeability measurements are made with a single fluid filling the pore space. This
is seldom the case in the reservoir situation except in water zones. Generally, two and
sometimes three phases are present, i.e. oil, water, and occasionally gas as well.
Here one would expect the permeability to either fluid to be lower than that for the single fluid
since it occupies only part of the pore space and may also be affected by interaction with other
phases. The concept used to address this situation is called relative permeability. The relative
permeability to oil, Kro, is defined as:

K ro =

K eo effective oil permeability


=
K
base permeability

(10.1)

K ew effective water permeability


=
K
base permeability

(10.2)

Similarly we can define:

K rw =
K rg =

K eg
K

effective gas permeability


base permeability

(10.3)

The choice of base permeability is not, in itself, critical provided it is consistently applied.
Conversion from one base to another is a matter of simple arithmetic. However,
experimentally, the base permeability is usually chosen as that measured at the beginning of
an experiment. For example, an experiment may start by measuring the permeability to oil in
the presence of an irreducible water saturation in the core. Water is then injected into the core,
and the oil permeability and water permeabilities measured as water replaces oil within the
core. The base permeability chosen here, would most commonly be the initial permeability to
oil at Swi.
Laboratory measurements are made by displacing one phase with another (unsteady state tests
see Figure 10.1) or simultaneous flow of two phases (steady state tests Figure 10.2). The
effective permeabilities thus measured over a range of fluid saturations enable relative
permeability curves to be constructed. Figure 10.3 shows an example of such a curve from an
unsteady state waterflood experiment. At the beginning of the experiment, the core is
saturated with 80% oil, and there is an irreducible water saturation of 20% due to the water
wet nature of this particular example. Point A represents the permeability of oil under these
conditions. Note that it is equal to unity, because this measurement has been taken as the base
permeability. Point B represents the beginning water permeability. Note that it is equal to zero
because irreducible water is, by definition, immobile. Water is then injected into the core at
one end at a constant rate. The volume of the emerging fluids (oil and water) are measured at

Dr. Paul Glover

Page 104

Formation Evaluation MSc Course Notes

Relative Permeability

the other end of the core, and the differential pressure across the core is also measured. During
this process the permeability to oil reduces to zero along the curve ACD, and the permeability
to water increases along the curve BCE. Note that there is no further production of oil from
the sample after Kro=0 at point D, and so point D occurs at the irreducible oil saturation, Sor.
Note also that Kro + Krw 1 always.

Dr. Paul Glover

Page 105

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 106

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 107

Formation Evaluation MSc Course Notes

Relative Permeability

It must be stressed, however, that these curves are not a unique function of saturation, but are
also dependent upon fluid distribution. Thus the data obtained can be influenced by saturation
history and flow rate. The choice of test method should be made with due regard for reservoir
saturation history, rock and fluid properties. The wetting characteristics are particularly
important. Test plugs should either, be of similar wetting characteristics to the reservoir state,
or their wetting characteristics be known so that data can be assessed properly.

Rigs for relperm measurement are often varied in design depending upon the exact
circumstances. Figure 10.4 shows an example of a typical rig piping diagram. The fluid flow
lines would be nylon or PTFE tube for ambient condition measurements (fluid pressures up to
a few hundred psi, and confining pressures up to 1500 psi), and stainless steel for reservoir
condition measurements (fluid pressures of thousands of psi, confining pressures up to 10,000
psi, and temperatures up to 200oC). These latter experiments are extremely complex, timeconsuming, and expensive especially if live fluids are to be used. The mean saturation in the
core is measured by collecting and measuring the volume of time-spaced aliquots of the
evolving fluids. However, there are various successful methods of monitoring the saturation
of the various fluids inside the core during the experiments. These are:

Dr. Paul Glover

Page 108

Formation Evaluation MSc Course Notes

Relative Permeability

(i) GASM (Gamma Attenuation Saturation Monitoring). Commonly used by BP, this uses
doped oil or water phases to attenuate the energy of gamma rays that travel through the core
perpendicular to the flood front. Each gamma source/detector pair measures the instantaneous
water and oil, or gas and oil saturation averaged over a thin cross section of the core. Up to 8
pairs are used to track the fluid saturations in the core during an experiment, giving a limited
resolution. Modern techniques use a single automated motorised source/detector pair.
(ii) X-Radiometry. Commonly used by US companies. It is similar to GASM, but uses xrays instead of gamma rays.
(iii) CT Scanning. Uses x-rays and tomographic techniques to give a full 3D image of the
fluid saturations in the core during an experiment. The spatial resolution is about 0.5 mm, but
is extremely expensive, and measurements can be made only every 5 minutes or so.
(iv) NMR Scanning. A very new application that is similar to the CT scanning. It has an
increased resolution, but is even more expensive.
The first two methods are commonly used, whereas the last two are rarely used due to their
cost.

10.2

Oil-brine relative Permeability Theory

Three cases will be considered:


(i) Water-wet systems
(ii) Oil-wet systems, and
(iii) The intermediate wettability case.
It should be remembered that in water-wet systems capillary forces assist water to enter
pores, whereas in the oil wet case they tend to prevent water entering pores.
Many reservoir systems fall between the two extremes, which does nothing to make laboratory
water-flood data easier to interpret. However, a knowledge of the two extreme cases allows
misinterpretation of intermediate data to be minimised.
Consideration must be given to flow rates. Close to the well bore, advance rates will be high,
further away, rates can be very low. This can be modelled in laboratory tests; but in the case
of oil wet systems, there is a tendency for low recoveries to be predicted due to end effects,
i.e. retention of wetting phase at test plug outlet face.

10.2.1 Water Wet Systems


Consider a water-wet pore system at Swi (generally 15 to 30%) some distance from well bore
such that flow rates are low, typically advancing at 1 ft/day. This is equivalent to about 4
cc/hr in a typical laboratory waterflood. The following sequence occurs as water migrates into
the rock:

Dr. Paul Glover

Page 109

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 110

Formation Evaluation MSc Course Notes

Relative Permeability

(i)

Figure 10.5. Initially at Swi, water is the wetting phase and will not flow. Kro = 1 and
Krw = 0.

(ii)

Figure 10.6. Water migrates in a piston like fashion, tending to displace most of the oil
ahead of it.

(iii) Figure 10.7. As water saturation increases oil flow tends to cease abruptly, and Sor is
reached.
(iv) Figure 10.8. Dramatically increasing the water flow rate (bump) has very little effect on
oil production or Krw. This is because capillary forces provide most of the energy
required for displacement of the oil.
If floods are carried out at too high a flow rate on water-wet cores the trapping mechanisms
present in the reservoir are not allowed to occur. Instead of entering small pores preferentially
by capillary forces, the water flows at a relatively higher velocity through larger pores, thus
tending to bypass groups of smaller pores containing oil. The Sor value obtained may then
differ from the true reservoir situation.

Dr. Paul Glover

Page 111

Formation Evaluation MSc Course Notes

Relative Permeability

Water wet systems are usually adequately described by low rate floods, and do not exhibit end
effects to any significant extent. Water wet data are characterised by:
(i) Limited oil production after water breakthrough.
(ii) Generally good recoveries.
(iii) Low Krw values at Sor.
Some typical data are presented in Figures 10.9 and 10.10. Points to take note of are the
limited amount of incremental data obtained (although this may be extended by using viscous
oils). This is caused by the rapid rise in water cut and the very short period of two phase flow
typical of water wet systems.

10.2.2 Oil Wet Systems


Consider water entering an oil-wet pore system containing (typically) very low water
saturations. The sequence of events from Swi is illustrated by Figures 10.11 to 10.14 as
follows:

Dr. Paul Glover

Page 112

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 113

Formation Evaluation MSc Course Notes

Relative Permeability

(i)

Figure 10.11. Capillary pressure considerations indicate that an applied pressure


differential will be required before water will enter the largest pore. The actual pressure
differential required is dictated by Eq. (8.1).

(ii)

Figure 10.12. Water flows through the largest flow channels first, Kro falls and Krw rises
rapidly.

(iii) Figure 10.13. After large volumes of water have flowed through the system, Sor is
reached. This equilibrium is attained slowly giving the characteristic prolonged slow
production of oil after early water breakthrough.
If waterfloods on oil wet core are carried out at too low a flow rate there may be inappropriate
retention of oil at the outlet face of the test plug. This is illustrated in Figure 10.14. At the
end of a low rate flood, Krw and the amount of oil produced are relatively low. If the flow rate
(and hence the pressure differential) are increased at this stage, substantial further oil
production occurs and Krw increases significantly. This situation does not model processes
occurring in the reservoir and should be avoided by appropriate choice of waterflood rate at
the beginning of the experiment.

Typical high rate valid oil wet data are shown in Figure 10.15 and 10.16.

Dr. Paul Glover

Page 114

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 115

Formation Evaluation MSc Course Notes

Relative Permeability

10.2.3 Intermediate Systems


Data from intermediate wetting rock-fluid systems can be difficult to assess, especially if a
single test mode has been used to obtain data. It is usually necessary to carry out a variety of
flood modes to fully assess end effects and rate dependence. A low rate flood followed by a
high rate bump flood will usually give an indication of the extent of end effects. Flow
reversal may indicate whether low recoveries are due to pore scale end effects evenly
distributed within the test plug. Steady state tests may be necessary to fully define the shape
of the relperm curves.
Typical intermediate wettability relperm data are shown in Figures 10.17 and 10.18. The
shape of the relperm curves is significantly different for the high and low rate floods.
However, the volume of oil produced is similar.

Figure 10.19 shows steady state data obtained from a core containing mobile kaolinite fines.
These were mobilised during the prolonged simultaneous flow of oil and brine during the
steady state test sequence. They have caused the water relative permeability to be suppressed.
Figures 10.17, 10.18 and 10.19 contain data obtained on the same test plug and illustrate the
need for more than one test made in obtaining valid relative permeability data. Note also that

Dr. Paul Glover

Page 116

Formation Evaluation MSc Course Notes

Relative Permeability

the data points are more reasonably spaced and are less scattered for the steady state test, but
there are fewer of them. The steady state test is more controlled because it takes much longer
to carry out, but the length of time required to come to equilibrium at each flow rate ratio (at
least a few days compared to less than one day for a whole unsteady state test) results in fewer
data points being taken.

Dr. Paul Glover

Page 117

Formation Evaluation MSc Course Notes

Relative Permeability

10.2.4 Relative Permeability Calculations


This section includes:
(i)

Calculation of relative permeability from laboratory waterflood data and the basic
equations from Johnson Bossler & Nauman (JBN).
(ii) Prediction of fractional flow from relperm curves and capillary pressure data.
(iii) Fractional flow from transition zones.
(iv) Variation of fractional flow with viscosity ratio.
I.

Calculation of Relative Permeability from Waterflood Experiments: JBN Analysis.

The experimental data generally recorded includes:


Qi
p
pi
Qo
Qw

=
=
=
=
=

Dr. Paul Glover

Quantity of displacing phase injected


Pressure differential
Pressure differential at initial conditions
Volume of oil produced
Volume of water produced

Page 118

Formation Evaluation MSc Course Notes

Relative Permeability

These data are analysed by the technique described by Johnson, Bossler and Nauman (see
reference list [4]), which is summarised below. Three calculation stages are involved:
(a)
(b)
(c)

The ratio Kro/Krw.


The values of Kro and hence Krw.
The value of Sw.

The method is aimed at giving the required values at the outlet face of the core which is
essentially where volumetric flow observations are made.
(a)

Kro/KrW The average water saturation (Swav)is plotted against Qi:

It can be shown that the fractional flow of oil, at the core outlet is given by:

fout =

d S wav
d Qi

(10.4)

Together with:

fout =

(b)

1
K
1 + rw o
K ro w

(10.5)

Kro A plot of p/pi against Qi is used to obtain the injectivity ratio IR:

Dr. Paul Glover

Page 119

Formation Evaluation MSc Course Notes

IR =

p i 1
p Q i

Relative Permeability

(10.6)

Kro is obtained by plotting 1/QiIR versus 1/Qi;

Dr. Paul Glover

Page 120

Formation Evaluation MSc Course Notes

Relative Permeability

and using the relationship;

1
K ro = fout
d (1 / Q i I R )
d (1 / Q i )

(10.7)

Knowing Kro/Krw from (a) above, then Krw can be calculated.


(c)

Using Welges correction to convert average saturations to outlet face saturations


(Swout);

Swout = Swav fout Q i

(10.8)

Thus Kro and Krw can be plotted against Swout to give the normal relative permeability curves.
II.

Prediction of Fractional Flow

Fractional flow can be predicted from capillary pressure data and relperm curves. Capillary
pressure data gives the saturations expected, and the relperm curves provide the values for Krw
and Kro at that saturation. Water cut can then be calculated.
Water and oil cuts are defined as follows:

Water Cut % =

Water production
x 100%
Total production

(10.9)

and

OilCut % =

Oil production
x 100%
Total production

(10.10)

Using the radial flow equations;

Qo =

2 K eo H t Pd
R
o ln e
Rw

(10.11)

and;

Dr. Paul Glover

Page 121

Formation Evaluation MSc Course Notes

Qw =

where;

Ht
Pd

=
=

2 K ew H t Pd
R
w ln e
Rw

Relative Permeability

(10.12)

Thickness of the producing zone being considered.


Drawdown pressure.

But from equations 10.1 and 10.2;

K eo = K ro K

(10.13)

K ew = K rw K

(10.14)

and

Thus, since the fractional water cut, fw, is defined as;

fw =

Qw
Qo + Q w

(10.15)

we can say:

fw =

III.

1
K
1 + w ro
o K rw

(10.16)

Fractional Flow from Transition Zones

Transition zones or zones with Sw at some value greater than Swi may present problems with
unsteady state tests. It may not be possible to perform an unsteady state waterflood starting at
Sw values greater than Swi, i.e. where the initial oil saturation is lower than the irreducible
saturation attained at infinite capillary pressure. The steady state test may be more
applicable in such cases. This situation frequently exists in transition zones before production
is started. When production commences the oil/water flow ratio should correlate with steady
state water drainage test data, i.e. carried out with Sw increasing. This is the most probable
direction in which saturation last changed to place the reservoir in its discovery state.
IV. Variation of Fractional Flow with Viscosity Ratio
For cases where capillary forces are negligible, it can be shown that the fractional flow of
water increases as the viscosity of water decreases relative to the oil viscosity. Using the term
mobility, defined as:

M ro for oil phase = K ro / o


Dr. Paul Glover

(10.17)

Page 122

Formation Evaluation MSc Course Notes

Relative Permeability

M rw for water phase = K rw / w

(10.18)

the Mobility Ratio = M ro / M rw

(10.19)

Eq. (10.16) is then expressed as:

fw =

1
1+

(10.20)

Thus fw decreases as w is increased or as o is decreased. The effect is shown in Figure 10.20.

Dr. Paul Glover

Page 123

Formation Evaluation MSc Course Notes

Relative Permeability

10.3 Laboratory Tests Available


There are at least ten usual variations of room condition tests, and each can also be done at
full reservoir conditions of confining pressure, fluid pressure, and temperature with live fluids
(Table 10.1).
Table 10.1 Common Laboratory Relperm Tests
Type
Oil/brine imbibition
Oil/brine drainage
Oil/brine imbibition
Oil/brine drainage
Gas/brine drainage
Gas/brine imbibition
Gas/oil drainage
Gas/oil imbibition
Gas/oil drainage
Gas/oil imbibition

Mode

Sw

So

Sg

Steady state
Steady state
Unsteady state
Unsteady state
Unsteady state
Unsteady state
Unsteady state
Unsteady state
Unsteady state
Unsteady state

Increasing
Decreasing
Increasing
Decreasing
Decreasing
Increasing
Sw=0
Sw=0
Sw=Swi
Sw=Swi

Decreasing
Increasing
Decreasing
Increasing
So=0
So=0
Decreasing
Increasing
Decreasing
Increasing

Sg=0
Sg=0
Sg=0
Sg=0
Increasing
Decreasing
Increasing
Decreasing
Increasing
Decreasing

The most representative and costly test is the reservoir condition waterflood. This is carried
out on core which has been restored to full reservoir conditions of temperature, overburden
loading, fluid contents (live crude) and wettability. Limited numbers of these tests are
performed to assess more economical room condition waterflood data.
In view of the large number of possibilities, detailed discussion here will be limited to those
most frequently studied, i.e. water-floods, steady and unsteady state, gas/brine drainage and
imbibition, and gas/oil drainage and imbibition.

10.3.1 Oil-Brine Relative Permeability


This is the most frequently requested relative permeability test. It attempts to simulate the
displacement of oil by a rising oil-water contact or a waterflood. The choices of test mode
available are unsteady state or steady state, and each has its limitations and advantages. In
general terms unsteady state tests are less time consuming than steady state tests, but can
suffer from uneven saturation distributions (end effects). Displacement rates can be modified
to accommodate wettability characteristics to some extent, and to model reservoir flow rates.
Steady state tests can be set up to avoid end effects but are more time consuming, requiring
time to reach equilibrium at each chosen oil/brine flow ratio.

10.3.2 Oil-Brine Unsteady State Test Procedure


Cleaned cores at Swi are confined in a Hassler or other type of core holder fitted with a
breakthrough detector and subjected to a constant brine flow. Data recorded are incremental
oil and brine production (in calibrated vials), the pressure differential across the core, and the
brine breakthrough point. The data are used to calculate the relative permeability
characteristics by using the Johnson, Bossler and Naumans technique.

Dr. Paul Glover

Page 124

Formation Evaluation MSc Course Notes

Relative Permeability

The normal full test sequence is as follows:


(i)

Miscibly clean core by flushing alternately with toluene and methanol; measure weight
saturated with methanol.
(ii) Saturate with formation brine without drying; measure weight saturated with brine.
(iii) Measure Kew at Sw=1.
(iv) Flood down to Swi at a suitable differential pressure.
(v) Measure Keo at Swi.
(vi) Carry out waterflood, recording pressure differential, incremental oil and water
production, etc. (data required for JBN analysis).
(vii) Use JBN analysis to calculate Keo, Kew, Kro, and Krw for various Swout and Swav.
(viii) Measure Kew , and calculate Krw at Sor before and after bump.
(ix) Clean, dry, measure KL and .
Flooding down to Swi is carried out in a Hassler or other type of core holder fitted with a
capillary pressure disc. This process may take several weeks, but has the advantage over
centrifuge techniques that even saturation distributions are obtained. Oil wet and intermediate
systems tend to flood to typically low values of Swi more rapidly, and at lower pressure
differentials than water wet systems. Figures 10.9, 10.15 and 10.17 show example data for
water-, oil- and intermediate wet cores.

10.3.3 Oil-Brine Steady State Test Procedure


These differ from the unsteady state tests in that oil and brine are flowed simultaneously
through the test plug at a fixed ratio until equilibrium is attained, Figure 10.2 (constant
pressure differential). The saturations were traditionally determined by demounting the plug
and weighing, but are now done using one of the methods discussed at the end of section 10.1.
The process is repeated with various oil/brine ratios, changing to suit the expected reservoir
history, to build up complete relative permeability curves (e.g., Figure 10.19). The effective
permeabilities are simply calculated using Darcys Law.
The disadvantages of steady state tests are that they are more time consuming both in manhours and elapsed time than unsteady state floods. It usually takes at least 24 hours for each
flow ratio to equilibrate, but this can extend to 72 hours for low permeability samples or
samples made from several core plugs abutted to each other to form a long test sample.
Estimation of saturation can be difficult for friable samples if grain loss occurs each time the
plug is removed for weighing. The methods of measuring fluid saturations in situ overcome
this problem.
Steady state tests have the advantage that end effects (which can affect certain unsteady state
tests) are eliminated. The test core is mounted between mixer heads made from adjacent core
material. These have similar wetting characteristics to the test plug and allow the correct flow
regime to fully establish itself before the test plug is entered. End effects then occur in the
outlet end piece instead of the test plug.

Dr. Paul Glover

Page 125

Formation Evaluation MSc Course Notes

Relative Permeability

10.3.4 Gas-Brine Relative Permeability Tests


Unsteady state tests are most common due to difficulties handling injection of gas over long
periods in steady state tests. Consequently, this section will only deal with unsteady state gasbrine relperm tests. The drainage cycle, i.e. gas displacing brine, models gas injection into a
brine saturated zone. Full relative permeability curves are generated and more importantly,
gas permeability at irreducible brine saturation. The imbibition cycle models movement of a
gas/water contact into the gas zone. Imbibition tests cannot be set up to give the full relative
permeability curves, but do give brine permeability at trapped gas saturation and the actual
trapped gas saturation itself. Typical gas-brine drainage and imbibition data are shown in
Figures 10.21 and 10.22.

The drainage test is performed by flowing gas (saturated with water vapour to ensure that the
gas does not evaporate the brine) into a brine saturated plug. Incremental gas and brine
production and pressure differential are recorded. Relative permeability curves can then be
calculated using JBN analysis.

Dr. Paul Glover

Page 126

Formation Evaluation MSc Course Notes

Relative Permeability

Figure 10.22 Gas/Brine Relative Permeability Data


Plug
Code

3
5
12

KL,
mD

12.8
411
103

Drainage

Imbibition

Kew at
Sw=100
% mD

Keo at
Swi,
mD

Kro at
Swi,
mD

Sw at
End of
Test

Sg at
end of
Test

Kew at
Sgi, mD

Krw at
Sgi, mD

Sg at end
of test

Sw at
end of
test

4.3
165
50

5.2
87
44

1.21
0.53
0.88

0.38
0.42
0.21

0.62
0.58
0.79

0.21
1.48
1.05

0.05
0.01
0.02

0.53
0.19
0.48

0.47
0.81
0.52

Imbibition data are obtained by recording the pressure differential across the core as brine is
flowed into the test plug initially at S=Swi+Sg. As the initially dominant gas phase is replaced
by more viscous water, the pressure differential rapidly increases to a maximum. It then falls
slowly as gas dissolves in the flowing brine. This dissolution is unavoidable to some extent,
but can be reduced by equilibrating the injected brine with the gas at pressure prior to
injection. It should be noted that the injected brine will not completely displace the gas, and a
trapped gas saturation will always remain. The maximum pressure differential is recorded and
used to calculate Krw at residual (trapped) gas saturation. Krw at trapped gas saturation can be
surprisingly low, values of 0.02 to 0.1 being frequently recorded.
It is interesting to consider the reservoir situation which is slightly, but significantly, different
from the laboratory technique. In the reservoir water migrates into the gas zone as pressure
declines, but unlike the core test, the gas saturation does not necessary decline. It tends to
remain high or increase slightly, since the trapped gas expands as pressure falls. This
maintenance or even increase in Sg tends to keep Krw low or reduce it even further. This
scenario operates in many reservoirs even if some of the gas migrates onwards and upwards.

10.3.5 Gas-Oil Relative Permeability Tests


Unsteady state tests can be performed in both drainage and imbibition modes. The drainage
mode (gas displacing oil) models gas advance into an oil zone, and usually yields full relperm
curves. The imbibition cycle provides data for an oil zone advancing into a gas cap but only
end point permeability and trapped gas saturation are obtained. It is worth considering the
mechanism occurring as an oil reservoir is depleted to a pressure below its bubble point. The
process which occurs is represented in Figures 10.23, 10.24 & 10.25.
(i)

Referring to Figure 10.23. Initially gas forms in discrete, immobile bubbles, which
reduces Kro very significantly.
(ii) Figure 10.24. As pressure falls further, the gas saturation increases. The bubbles
eventually become connected and give rise to a significant gas permeability. The
saturation at which gas becomes mobile is termed the critical gas saturation. Krg rapidly
increases and Kro further declines. The relative flow rate of oil is further reduced by the
lower viscosity and higher mobility of the gas.
(iii) Figure 10.25. Eventually the oil droplets become discontinuous and only gas is
produced.

Dr. Paul Glover

Page 127

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 128

Formation Evaluation MSc Course Notes

Dr. Paul Glover

Relative Permeability

Page 129

Formation Evaluation MSc Course Notes

Relative Permeability

Laboratory gas-oil relative permeability tests are performed in a similar manner to the gasbrine tests. If required, the tests can be performed with connate water present, but this
requires that brine saturated cores be flooded to Swi with oil prior to gas flooding. The relative
merits of tests with and without connate water have not yet been fully investigated. It can be
argued that the connate water will be immobile and this has been found to be true in some
experiments. However, where connate water is present we have noticed that Krg tends to
show a more concave upwards curve than when it is absent. The situation is very complex,
but could possibly be affected by the wetting characteristics of the rock. The effect is shown
in Figures 10.26 & 10.27.

10.3.6 Relative Permeability Data Treatment


Interpretation and use of relative permeability data to predict individual well or reservoir
performance can be complicated by lateral variations in rock properties. Thus, although the
laboratory tests can adequately describe the behaviour of a particular test plug, modelling of a
well or reservoir performance may require modified relative permeability data. Correlations
of overall curve shape, cross-over points, recovery at a given produced volume, brine
permeability at residual oil saturation etc., must all be made with reference to lithology,
permeability, and initial fluid saturations.
Choice of test method will be governed by application of data, i.e. high flow rate for near the
well bore, and low flow rates away from the well bore. As stressed previously, no one test
method can fully describe a system and choice of data will be influenced by laboratory scale
limitations; in particular, end effects in oil wet cores and problems sometimes caused by
wettability alterations and mobile fines.
Good petroleum engineering reports should highlight any experimental difficulties
encountered and indicate the most reliable data. However, it is often impossible to assess
service company data since flow rates are seldom constant and p/pi versus Qi curves are not
reported. If you are ever in a position to commission this type of work, ensure that provision
of this information is part of the contract.

Dr. Paul Glover

Page 130

Formation Evaluation MSc Course Notes

Commissioning Studies

Chapter 11: Commissioning Studies


11.1 Introduction
It is not easy to specify RCAL and SCAL programmes on a general basis. Each reservoir (or
well) has to be considered on its own merits. The numbers of each particular test required
will depend on the permeability/porosity distribution. The range of tests required will depend
on previous experience with the particular reservoir rock/fluid combination under study.
Where previous experience is available and there is good agreement between routine core
analysis, log and well performance data, the numbers of tests and possibly scope can be
reduced. Where anomalies exist it will be advantageous to increase coverage of some
parameters and possibly introduce special test sequences to assess the more unusual core and
fluid interactions connected with formation damage.
The more common reservoir situations encountered are summarised below. In each case it is
the SCAL that has been concentrated upon. Routine core analysis commissioning is usually
done by internal protocol. For example, plugging every foot wherever the strength of the core
fabric allows it, basic chemical cleaning, and then He porosity and gas permeability on the
core plugs. Sometimes Hg porisimetry, CT scanning and petrographic analyses are also
carried out on a small number of plugs.
When commissioning a SCAL study the FIRST step is to review ALL the RCAL data that
has already been obtained. The RCAL information effectively provides a pre-study that will
help identify possible problems in carrying out the SCAL work. These problems may include
(i) swelling clays, (ii) friable and unconsolidated core, (iii) drilling fluid contamination, (iv)
mobile fines, etc.
Also, it is important to seek the technical advice and expertise of:
v Anyone with experience of the rock properties for the field in question.
v The technical staff and managers carrying out the work.
v Any other individual or source that may provide data that could help save expensive
mistakes.
Remember time is money, but if corners are cut such that the real requirements are not well
defined, or data that you already have is not used, or possible problems with the rocks are not
recognised, then more money will be wasted. SCAL data is very expensive to collect. Hence,
ask yourself the following points during SCAL commissioning:
1.
2.
3.
4.
5.

What data do you have already from RCAL?


What other knowledge is there that you can draw upon?
What is the type of field, drive and hydrocarbon?
What do you want to do with the SCAL data?
Will your projected SCAL campaign provide sufficient data for the purpose it was
designed - Is it fit for purpose?
6. Is there any information that would be missing? If it is required, commission it!
7. Is there a specified test that will not tell you any further information concerning the field?
If there is drop it!

Dr. Paul Glover

Page 131

Formation Evaluation MSc Course Notes

Commissioning Studies

8. Have you commissioned sufficient cores of each type of test to provide a reliable and
representative coverage of the well section of interest? If you do not know how to judge
this, get advice!
9. Does the data you have flag any possible problems with the particular rock to be analysed?
If there are problems you must talk to the people carrying out the tests. They should be
warned of problems, and in most cases will be able to advise you on the best technical
approach to either overcome, avoid or minimise the problem.
10. Does the service company know your needs sufficiently well to provide good and
informed service? Good communications leads to better data!
11. Have you asked the service company managers and technical staff for their expert advice?
12. Do you know the timescale for the scheduled work, and is it fixed? If not fix it SCAL
tests are of known duration and can be scheduled well.
13. Have you specified the errors that are acceptable on the measurements? To do this you
will need to communicate with the service company. Ensure that measurement errors,
which always occur, are sufficiently small or manageable that the data is fit for purpose.
14. Have you specified the data to be provided in a format that will the of best use to you, i.e.,
on CD-ROM in Excel format as well as a written report?
A little time spent planning can save an enormous amount of money.

11.2 Dry Gas Reservoirs


These represent the simplest case but the relative importance of the suggested tests (Figure
11.1) will depend upon the nature of the reservoir. Particular care should be taken with highly
fractured reservoirs where gas/brine contacts move rapidly. Measurement of trapped gas
saturation is particularly important here. With sandstone reservoirs brine permeability in the
water zone may be significantly lower than indicated by routine measurements. This can be
readily assessed from a relatively small number of tests on preserved water zone cores.

Dr. Paul Glover

Page 132

Formation Evaluation MSc Course Notes

Commissioning Studies

Figure 11.1 Basic SCAL Programme for a Dry Gas Reservoir 100
Thick with a Single Facies Type
Test Sequence

1
2

4
5
6

Drill, trim and clean by flushing as many plugs as are


obtainable from preserved core (up to 60)
Capillary pressure (air-brine)
Resistivity index
Saturation exponent
Formation factor
Cementation factor
Formation factor at overburden pressure
Pore volume compressibility at overburden pressure
Brine permeability at overburden pressure
Ambient condition relative permeability gas floods
Drainage and imbibition cycles
Water zone permeabilities
Brine permeabilities of preserved core - plug samples
Clean by routine methods
Routine porosity (He method)
KL gas permeability
Plug description, thin section, and SEM studies to link
SCAL properties with sedimentological characteristics
Typical cost

Number of
Tests
Proposed

Typical
Charge per
Test (Manhours)

As required

10-20

15

10-20

12

10

30

As required

As required

700-1000 man-hours
(approximately 800,000)

11.3 Oil Reservoir without Gas Cap


The situation with oil reservoirs becomes more complex than a dry gas reservoir as:
(i)
(ii)

Transition zones are usually from a more significant portion of the reservoir.
Flow characteristics and relative permeabilities are strongly influenced by wettability.

The static tests, i.e. capillary pressure, resistivity index, formation factors etc., are basically
straightforward, (Figure 11.2). However, dynamic tests, i.e. relative permeability, are more
complicated, and choice of type and test mode depends upon wettability. Wettability
measurements should be considered as an essential preliminary to choice of relative
permeability test. If waterflooding is envisaged then wettability is extremely important, as
will be water zone brine permeabilities on preserved core.

Dr. Paul Glover

Page 133

Formation Evaluation MSc Course Notes

Commissioning Studies

Figure 11.2 Basic SCAL Programme for an Oil Reservoir 100 Thick
with a Single Facies Type

1
2

4
5
6

Test Sequence

Number of
Tests
Proposed

Typical
Charge per
Test (Manhours)

Drill, trim and clean by flushing as many plugs as are


obtainable from preserved core (up to 60)
Capillary pressure (air-brine)
Resistivity index
Saturation exponent
Formation factor
Cementation factor
Formation factor at overburden pressure
Pore volume compressibility at overburden pressure
Brine permeability at overburden pressure
Wettability tests on preserved and cleaned core
Ambient condition relative permeability tests
(mode chosen depending on wettability measurements)
Reservoir condition waterfloods. Possibly in preference
to (5 above) depending on the characteristics of the
core
Water zone permeabilities
Pore throat size distribution measurements by mercury
injection
Clean by routine methods
Routine porosity (He method)
KL gas permeability
Plug description, thin section, and SEM studies to link
SCAL properties with sedimentological characteristics
Typical cost

As required

10-20

15

10-20

12

6
10-15

15
30

Up to 6

120

10

As required

As required

700-1700 man-hours
(approximately 1,400,000)

11.4 Oil Reservoir with Gas Cap


The static tests are basically the same as for 11.2 above. Choice of dynamic relative
permeability tests will depend upon the expected movement of oil/water and gas/oil contacts.
If expansion of the gas cap into the oil zone is envisaged, gas-oil relative permeability at
connate water tests are desirable. Similarly if the reservoir is being waterflooded for pressure
maintenance or to reduce gas cap size; the imbibition gas/oil test will provide valuable data on
oil permeability at trapped gas saturation and the trapped gas saturation itself.

Dr. Paul Glover

Page 134

Formation Evaluation MSc Course Notes

Commissioning Studies

11.5 Gas-Condensate
The flow regimes and saturation changes which occur in condensate reservoirs are among the
most difficult to model in the laboratory, and are extremely rare.

11.6 Formation Damage


The term formation damage generally describes permeability reduction brought about by:
(i)
(ii)
(iii)
(iv)

Movement of fines.
Introduction of particulate matter.
Introduction of incompatible fluids.
Introduction of fluids upsetting desired relative permeability behaviour.

It is especially difficult to specify a general scheme of formation damage tests. The particular
reservoir fluids, minerals, saturation change directions, and introduced fluid compositions
should be considered when defining a programme. Two situations will therefore be briefly
covered which illustrate the means of damage detection and the applicability of single and two
phase tests. The cases considered here are poor and declining injectivity in a water injection
well, and formation damage caused by drilling muds.

11.6.1 Poor and Declining Injectivity


The possibilities here are:
(i) Particulate matter in injection water.
(ii) Mobile fines within reservoir rock.
(iii) Incompatible waters causing clay swelling.
These processes can be tested for by the following methods:
(i)

The problem of particulate matter in injection water should be taken care of by proper
filtration but could be tested for with on site core tests. The tests however tend to be
pessimistic and indicate greater permeability decline rates than are encountered
downhole.

(ii)

The presence of mobile fines can be detected fairly readily in the laboratory.
Permeability to liquids (brines) are observed and plotted against throughput. Changes
occur with throughput and flow direction when fines move to block pore throats, Figure
11.3.

(iii) The sensitivity of a formation to brine composition can be assessed by core throughput
tests with changing brine compositions, Figure 11.3. Simple clay swelling effects are
observed as reversible permeability changes. However, it is possible that some particles
become dislodged during the tests and then behave as mobile fines.

Dr. Paul Glover

Page 135

Formation Evaluation MSc Course Notes

Commissioning Studies

10.6.2 Drilling Mud Formation Damage


A recent study indicated that single phase (liquid) permeability tests cannot necessarily be
relied upon to predict formation damage for a two phase situation. Single phase tests
indicated that oil based mud filtrate permeability was greater than for water based mud filtrate,
implying permeability damage by the water based mud. However, when the mud filtrates
were displaced with gas, the effective gas permeabilities were the same in both instances. This
case has been simplified, as other factors, such as relative permeability, fluid saturations and
volume throughput required to achieve recovery of gas permeability; also need to be
considered when interpreting the permeability data.

Dr. Paul Glover

Page 136

Formation Evaluation MSc Course Notes

Abbreviations

Abbreviations
A
B
Cp
FF,F
E
Ht
Kbrine
Kg
KL
Ko
Keo
Kr
Kro
Krw
KSFW
Kw
Kew
l ,L
m
m*
Mr
n
n*
P, p
Pc
Pd
Pm
Qo
Qi
Qv
Qw
Re
RCAL
Relperm
Ro
Rw
Rt
SCAL
So
Sor
Sgt
Sw
Swi
t

Dr. Paul Glover

Cross sectional area


Specific Counterion Activity (Waxman-Smits)
Capillary Pressure
Formation factor
Tortuosity factor
Producing zone thickness
Brine permeability
Gas permeability
Equivalent liquid permeability (Klinkenberg corrected gas permeability)
Oil permeability
Effective oil permeability
Relative permeability
Relative permeability to oil
Relative permeability brine
Permeability to simulated formation water
Brine/water permeability
Effective brine/water permeability
Length
Cementation factor
Cementation factor (corrected)
Mobility
Saturation exponent
Saturation exponent (corrected)
Pressure
Capillary pressure (psi)
Drawdown pressure
Mean flowing pressure
Volume oil produced
Volume water injected
Cation exchange capacity meq/ml
Volume water produced
Effective reservoir radius
Routine core analysis
Relative Permeability
Core resistivity
Brine resistivity (or wellbore diameter)
Core resistivity at reduced Sw
Special Core Analysis
Oil saturation
Residual oil saturation
Residual trapped gas saturation
Brine saturation
Initial brine saturation
Time (secs)

Page i

Formation Evaluation MSc Course Notes

Abbreviations

Abbreviations continued
V, v

Dr. Paul Glover

Volume
Interfacial tension
Mobility ratio
Contact angle
Porosity
Oil density
Brine density
Oil viscosity
Water (brine) viscosity

Page ii

Formation Evaluation MSc Course Notes

Selected References

Selected References
[1]

Permeability Characteristics of Magnus Reservoir Rock. Heaviside, Langley and Pallatt, 8th Formation
Evaluation Symposium Trans. Mar. 1983. London.

[2]

Errors in Laboratory Measurements of Formation Resistivity Factor, by A.E. Worthington. S.P.W.L.A.


16th Annual Logging Symposium 4-7 June 1975.

[3]

Comments on obtaining Accurate Electrical Properties of Cores, by Hoyer, S Spann. S.P.W.L.A. 16th
Annual Logging Symposium 4-7, June 1975.

[4]

Calculation of Relative Permeability from Displacement Experiments, Johnson, Bossler and Nauman, Pet.
Trans. AIME (1959), 216, p.370.

[5]

EHRLICH, R., CRABTREE, S.J., HORKOWITZ, K.O. & HORKOWITZ, J.P. 1991. Petrography and
reservoir physics 1: objective classification of reservoir porosity. The American Association of Petroleum
Geologists Bulletin, 75, 1547-1562.
Abstract: Porosity observed in thin section can be objectively classified using a combination of digital
acquisition procedures and pattern recognition algorithms. Pore types are derived from the frequency
distributions of sizes and shapes of patches of porosity exposed in thin section. Each pore type is
represented by a characteristic distribution of sizes and shapes found in thin section. Most sandstone
reservoirs contain fewer than six pore types. Much of the variability in reservoir physics is associated with
changes in pore type abundance. The advantages of this approach to porosity classification are (1) the
criteria for classification are objectively defined, (2) classification procedure is rapid, accurate, and
precise, (3) pore types are understood easily in terms of conventional genetic classification schemes, and
(4) pore type data are related strongly to petrophysical properties.

[6]

MCCREESH, C.A., EHRLICH, R. & CRABTREE, S.J. 1991. Petrography and reservoir physics 2:
relating thin section porosity to capillary pressure, the association between pore types and throat size. The
American Association of Petroleum Geologists Bulletin, 75, 1563-1578.
Abstract: Porosity in reservoir rocks is configured into a few types of pores whose size and shape are
controlled by depositional fabric and processes. The size, shape, and abundance of each pore type can be
objectively determined from thin section using image analysis and pattern recognition procedures. Each
pore type tends to be associated with a limited range of throat sizes. The association between pore type
and throat size can be determined using regression procedures linking pore type data obtained from thin
section with capillary pressure data. To do so, a set of samples is required wherein the association between
pore type and throat size is fixed, but where pore type proportions vary between samples. This condition
is met by a sample suite representing reservoir facies from a single core or, in many cases, from a single
field. The relationship between pore type and throat size is an effective means to relate reservoirs in terms
of the efficiency of the porous to multiphase flow. Parameters derived from the relationship can be used to
construct accurate physical models that subdivide physical response in terms of the contributions of each
pore type.

[7]

EHRLICH, R., ETRIS, E.L., BRUMFIELD, D., YUAN, L.P. & CRABTREE, S.J. 1991. Petrography
and reservoir physics 3: physical models for permeability and formation factor. The American
Association of Petroleum Geologists Bulletin, 75, 1579-1592.
Abstract: Permeability and formation factor are physical properties of porous rocks useful for assessing
reservoirs. Neither property varies consistently as porosity varies. The relationship of both properties to
porosity is complex, being sensitive to the structure of the porous microstructure, i.e., the sizes of pore
throats, the numbers and sizes of pores, and the relationships between pores and throats. Physical models
to account for these factors require parameters that describe physically relevant properties of the
microstructure. A partial characterization of the relationship between pores and throats is embodied in the
relationship between pore type and throat size. This relationship is derived by combining data obtained
from thin sections, from which pore types are derived via image analysis, and mercury injection
porosimetry, which quantifies throat size information. Parameters derived from such a combination are

Dr. Paul Glover

Page iii

Formation Evaluation MSc Course Notes

Selected References

sufficient to construct simple physical models for permeability and electrical conductivity (inverse
formation factor). These models assume a porous medium that has large numbers of flow paths parallel to
the potential gradient, such that flow has little tortuosity (i.e., flow parallel to bedding). The contributions
of each pore type to permeability and electrical conductivity are computed. Calculated values are close to
measurement values. A constant of proportionality is the same for all samples from a reservoir, but can
vary between reservoirs, is required, and must have values ranging (for sandstones) from about 2.5 to 3.5
for permeability and 5.0 to 7.0 for conductivity. These values are consistent for an efficiently packed
fabric. One result of such modeling is a physical model of Archie's cementation exponent m as the ratio of
the logarithms of the cross sectional throat area to pore area (per unit area).
[8]

WARDLAW, N.C. & TAYLOR, R.P. 1976. Mercury capillary pressure curves and the interpretation of
pore structure and capillary behaviour in reservoir rocks. Bulletin of Canadian Petroleum Geology, 24,
225-262.
Notes: A classic on this subject. Explains the reasons behind various aspects of injection and withdrawal
curves, by looking at SEM of rocks studied.

[9]

VAVRA, C.L., KALDI, J.G. & SNEIDER, R.M. 1992. Geological applications of capillary pressure: a
review. The American Association of Petroleum Geologists Bulletin, 76, 840-850.
Notes: Important discussion of interpreting mercury injection porosimetry results.

[10]

PITTMAN, E.D. 1992. Relationship of porosity and permeability to various parameters derived from
mercury injection-capillary pressure curves for sandstones. The American Association of Petroleum
Geologists Bulletin, 76, 191-198.
Notes: As mercury injection tests are expensive and not abundant, derives relationships using multiple
regression on large database of samples. Empirical equations make it possible to construct pore aperture
radius distribution curves from core analysis porosity and permeability.

[11]

BLIEFNICK, D.M. & KALDI, J.G. 1996. Pore geometry: control on reservoir properties, Walker Field,
Columbia and Lafayette counties, Arkansas. The American Association of Petroleum Geologists Bulletin,
80, 1027-1044.
Notes: An oolite carbonates sequence. Useful discussion on interpreting mercury injection porosimetery
results.

[12]

RINGROSE, P.S., SORBIE, K.S., FEGHI, F., PICKUP, G.E. & JENSEN, J.L. 1993. Relevant reservoir
characterisation: recovery process, geometry and scale. In Situ, 17, 55-82.
Notes: With miscible-gas flood, large-scale geometry may be more important than the internal small-scale
structure. With waterflood, small-scale structure likely to be dominant. Emphasises must think not only
about the rock but also the fluids and the recovery process.

[13]

CORBETT, P.W. & JENSEN, J.L. 1993. Quantification of variability in laminated sediments: a role for
the probe permeameter in improved reservoir characterization. In: NORTH, C.P. & PROSSER, D.J. (eds)
Characterization of fluvial and aeolian reservoirs. Geological Society special publication 73, London,
433-442.

[14]

HUANG, Y., RINGROSE, P.S. & SORBIE, K.S. 1995. Capillary trapping mechanisms in water-wet
laminated rocks. SPE Reservoir Engineering, 10, 287-292.
Abstract: Most floods in sandstone cores are performed either in almost homogeneous samples or else in
core samples of uncertain heterogeneity. As a result, the interaction of small-scale sedimentary
heterogeneity with the fluid mechanics of water-oil displacement cannot be adequately understood or
quantified. The results are reported from low-rate, drainage-imbibition floods in a 20 x 10 x 1 cm slab or
cross-laminated heterogeneous sandstone. The laminated aeolian sandstone was characterized by detailed

Dr. Paul Glover

Page iv

Formation Evaluation MSc Course Notes

Selected References

probe permeameter mapping prior to setting it in a resin cast. The distribution of porosity, permeability,
irreducible water, and residual oil saturation were subsequently monitored using CT scanning techniques.
The low-rate imbibition floods show that between 30 and 55% of original oil may be trapped in isolated
high permeability lamina. This work shows the importance of recognizing the role of core-scale
heterogeneity in the laboratory measurement of waterflood behavior, i.e., the interaction of capillary
forces with rock structure. The practice of performing high-rate floods on rock samples assumed to be
heterogeneous is unwise and can lead to erroneous conclusions. The work has major implications for (1)
2-phase petrophysical measurements, (2) the assessment of residual/remaining oil, and (3) multiphase flow
scaleup.
[15]

MCDOUGALL, S.R. & SORBIE, K.S. 1992. Network simulations of flow processes in strongly wetting
and mixed-wet porous media. In: CHRISTIE, M.A., DA SILVA, F.V., FARMER, C.L., GUILLON, O.,
HEINEMANN, Z.E., LEMONNIER, P., REGTIEN, J.M.M. & VAN SPRONSEN, E. (eds) ECMOR III:
Proceedings of the third European conference on the mathematics of oil recovery. Delft University Press,
Delft, Netherlands, 169-181.
Notes: Deriving 2-phase flow parameters such as relative permeability and capillary pressure from
microscopic considerations.

[16]

MCDOUGALL, S.R. & SORBIE, K.S. 1995. The impact of wettability on waterflooding: pore-scale
simulation. SPE Reservoir Engineering, 10, 208-213.

[17]

PICKUP, G.E., RINGROSE, P.S., JENSEN, J.L. & SORBIE, K.S. 1994.
sedimentary structures. Mathematical Geology, 26, 227-250.

Permeability tensors for

Abstract: Accurate modeling of fluid flow through sedimentary units is of great importance in assessing
the performance of both hydrocarbon reservoirs and aquifers. Most sedimentary rocks display structure
from the millimeter or centimeter scale upward. Flow simulation should therefore begin with grid blocks
of this size in order to calculate effective permeabilities for larger structures. Several flow models for
sandstones are investigated, and their impact on the calculation of effective permeability for single phase
flow is examined. Crossflow arises in some structures, in which case it may be necessary to use a tensor
representation of the effective permeability. Conditions are established under which tensors are required,
e.g., in crossbedded structures with a high bedding angle, high permeability contrast, and laminae of
comparable thickness. Cases where the off-diagonal terms can be neglected, such as in symmetrical
systems, are also illustrated. The method of calculating tensor permeabilities may be extended to model
multiphase flow in sedimentary structures.

Dr. Paul Glover

Page v

S-ar putea să vă placă și