Reservoir Engineering Notes 2018-2019 Revised
Reservoir Engineering Notes 2018-2019 Revised
Blunt
Reservoir Engineering
Martin J. Blunt
Lecture Notes
TABLE OF CONTENTS
Table of contents 2
1. Course overview and motivation 5
Syllabus 6
Course structure 6
Coursework 7
Recommended books 7
2. Philosophy 9
Contact information 9
3. Introduction to reservoir engineering 10
3.1 The three main concepts: Material balance, Darcy’s Law and data integration 10
3.2 What is a reservoir and what is a porous medium? 11
3.3 Fluid pressures 12
3.4 Oil initially in place 14
3.4.1 Definition of porosity and saturation 14
3.4.2 Conversion from reservoir to surface volumes 15
3.5 Oil production 17
3.6 The world’s largest oil fields 19
3.7 Fluid pressure regimes 24
3.8 Reservoir fluids 28
3.9 Phase behavior 34
4. Material balance 39
4.1 Material balance for gas reservoirs 41
4.1.1 Connate water and pore volume compressibility 44
4.1.2 Water drive gas reservoirs 47
4.1.3 Simple aquifer model 49
4.1.4 Aquifer fitting 50
4.1.5 Impact of residual gas and final recovery 53
4.2 Material balance for oil reservoirs 55
4.2.1 Production above the bubble point 60
4.2.2 Solution gas drive 63
4.2.3 Gas cap drive 66
4.2.4 Natural water drive 68
4.2.5 Compaction drive 69
4.2.6 Rate dependence 70
4.2.7 Recap on material balance 70
5. Decline curve analysis 72
5.1 Exponential decline 73
5.2 Hyperbolic decline 73
6. Multiple phases in equilibrium 76
6.1 Young-Laplace equation 76
6.2 Equilibrium at a line of contact 77
6.3 Spreading coefficient 78
6.4 Two fluids in a capillary tube 79
6.5 Wettability 81
6.5.1 Wettability alteration 81
6.5.2 Contact angle hystersisis 82
7. Porous media 86
7.1 X-ray imaging 86
7.2 Electron microscopy to image micro-porosity 90
7.3 Topologically representative networks 91
8. Primary drainage 94
8.1 Typical values of the capillary pressure 95
8.2 How is capillary pressure measured? 96
9. Imbibition 100
9.1 Pore-scale displacement, trapping of the non-wetting phase and snap-off 100
9.2 Pore-scale images of trapped phases 104
9.3 Typical capillary pressure curves and secondary drainage 110
9.4 Different displacement paths and trapping curves 111
10. Leverett J function 117
10.1 Capillary pressure and pore size distribution 120
11. Displacement processes in mixed-wet media 126
11.1 Oil layers 126
11.2 Effect of wettability on capillary pressure 128
11.2.1 Weakly water-wet media 128
11.2.2 Capillary pressures for mixed-wet media 128
11.2.3 Oil-wet systems 132
11.3 Trapping curves in mixed-wet systems 132
11.4 Transition zones 134
11.5 Amott wettability indices 135
11.5 Example exercises 136
12. Fluid flow and Darcy’s law 138
12.1 Stokes flow 138
12.2 Reynolds number and flow fields 139
12.3 Averaged behaviour and Darcy’s law 145
12.4 Other ways to write Darcy’s law and hydraulic conductivity 148
12.5 Units of permeability and the definition of the Darcy 149
12.6 Definition of flow speed and porosity 149
12.7 Estimating permeability 150
12.8 Example Problem in calculating permeability 152
13. Molecular diffusion and concentration 154
14. Conservation equation for single-phase flow 157
14.1 Analytical solution of the advection-diffusion equation 161
14.2 Diffusion and dispersion 164
15. Capillary and Bond numbers 170
16. Relative permeability 173
16.1 Relative permeabilities for sandstones and predictions using pore-scale modelling 176
16.1.1 Effect of wettability in sandstones 179
16.2 Imbibition and oil recovery processes 181
16.3 Analysis of relative permeability in mixed-wet carbonates 187
16.3.1 Pore structure and connectivity 187
16.2.3 Effect of fractional wettability on relative permeability 193
16.4 Comparison of network model results with experimental data 202
16.5 Impact of relative permeability on field-scale recovery 207
Access to sustainable, affordable water, energy and food are some of the major technological
challenges of the 21st century. How do we manage precious and scarce water resources, while
preventing pollution? How do we extract the most of our remaining supplies of conventional
oil and gas? How can we extract, safely, shale gas and oil? Can we collect and store carbon
dioxide in the subsurface to prevent atmospheric emissions and help avoid dangerous climate
change? Can we manage to provide sufficient energy for a growing world’s population with
an aspiration for improved prosperity? An understanding of these challenges involves
multiphase flow in porous media – the flow of water, oil and gas with associated pollutants –
underground in geological formations.
The subject of multiphase flow in porous media is undergoing a revolution, not just as a result
of its many important applications, but because of developments in our quantitative
understanding of how fluids are arranged and move, combined with the ability to image fluids
at the micron scale inside rocks.
This course will apply concepts of multiphase flow in porous media to understand and design
recovery from oil and gas reservoirs. This is one of the major challenges referred to above –
at present we recover only around one third of the oil from fields we have discovered. How
can we improve this to the 50-60% now achievable with the best engineering methods, and
beyond?
This course will describe how different methods can be used to:
assess the development potential of oil and gas reservoirs;
identify the principal displacement mechanisms controlling performance;
predict recovery and oil-in-place;
understand reservoir simulation methods.
It is assumed that you already know about hydrocarbon phase behaviour, reservoir simulation
and the principal reservoir drive mechanisms. You will also need to know Darcy’s law and the
meaning of relative permeability and fractional flow, although these are described again in
these notes.
The emphasis will be on learning fundamentals with time taken to cover basic concepts. I will
not repeat details that are well covered in other textbook, or which are not strictly relevant
to this course. Furthermore, these notes will not illustrate the concepts with field examples:
this is better left to project work or indeed industrial experience. This is not a manual for
reservoir engineers, but a teaching tool to establish the fundamentals.
SYLLABUS
COURSE STRUCTURE
The course will be presented in lectures and examined both as a written exam (80%) and as
an assessed exercise, which will be the 2018 examination on page 339 (20%). There will be
no oral examination. In addition, optional past examination papers can be attempted for
revision.
COURSEWORK
Tutorial questions are included at the end of these notes, as well as past examination papers.
These questions will be used to test your knowledge and understanding of the material
presented in lectures. Some of the questions will be done in class, others are for homework
and others are for revision. Answers to some of the questions, where appropriate, will be
given during the class. I will set a previous year’s examination as a homework exercise on
which I can provide feedback, as mentioned above.
RECOMMENDED BOOKS
This is a fabulous research reference book that covers much of the scientific material in these
notes and contains a lot of experimentally-based physical insight into multiphase flow.
9. Dynamics of Fluids in Porous Media, J. Bear, Dover Publications, Inc, New York, (1972).
A classic in its field – indeed it help establish the subject of flow in porous media as a discipline.
Very mathematical, but contains a lot of useful information.
10. Capillary and Wetting Phenomena: Drops, Bubbles, Pearls, Waves, P-G de Gennes, F.
Brochard-Wyart and D. Quéré, Springer (2002).
The first author is the charismatic, and now sadly deceased, physics Nobel Prize winner, Pierre-
Giles de Gennes. A fascinating book, packed full of interesting analysis, but not directly relevant
to flow in porous media. Acquire a French accent, light up a Gaulois, wave your hands and
voilá – brilliant insights into physics!
11. Multiphase flow in permeable media: a pore-scale perspective, M J Blunt, Cambridge University
Press (2017).
This is my own recently-published book that covers the material on multiphase flow in these
notes in more detail.
12. Reservoir Engineering, M J Blunt, Volume 2 of the Imperial College Lectures in Petroleum
Engineering, World Scientific Press (2017).
These are essentially these notes in book form: we have published a series of volumes based
on the lecture notes for the MSc course.
2. PHILOSOPHY
These notes contain a synopsis of the class and cover most of the material necessary to
understand the course. However, there are places where explanations and examples given in
lectures are essential to understand the material fully, so be prepared to make some
supplementary notes. I like to encourage discussion in lectures, so don’t be shy about asking
questions.
You will only be examined on material covered in lectures. Some of the reading assignments,
and large portions of the notes, deliberately cover extra material and are there mainly for your
own interest and to add extra background. This material will not be examined unless it repeats
material given in class. However, doing the reading will definitely help you understand the
class better and help with the homework problems.
CONTACT INFORMATION
I will be in my office when I am not teaching during the course, and I welcome questions, and
comments, either during the class or afterwards. My email address is
m.blunt@imperial.ac.uk.
The main application of this course is to the understanding of how oil, water and gas flow
deep underground with application to hydrocarbon recovery.
3.1 THE THREE MAIN CONCEPTS: MATERIAL BALANCE, DARCY’S LAW AND
DATA INTEGRATION
Before I present any details, there are three main points that need to be understood by any
good reservoir engineer. In the end, everything can be expressed with reference to one of
these three fundamental concepts.
1. Material balance. Mass is conserved: what leaves a reservoir (is produced) minus
what is injected is the change of mass in the subsurface. For every field, under
every circumstance, a reservoir engineer needs to check material balance – ideally
by hand – to understand and interpret production data. This will be the basic
principle on which I will base the analysis of fields under primary production.
Furthermore, it lies at the heart of the derivation of the flow equations used to
predict flow performance. But for this, we also need an equation for flow – point
2 below.
2. Darcy’s law for fluid flow. Fluid flows in response to a pressure gradient. The
linear relationship between the gradient of pressure (or, more generally, the
potential) and flow rate is Darcy’s law. It is the basis for any understanding and
prediction of flow.
3. Look at all the data and have a coherent, consistent understanding of the field.
A reservoir engineer assesses different information from several sources:
geological interpretations, seismic surveys, log analysis, core analysis and fluid
properties combined with production (rate and pressure) data. All of this data
needs to be incorporated into a model of the reservoir to predict future
performance and design production. A model in this context is not solely a
complicated computer realization of what the field might be like, but more a
conceptual understanding of the field that includes the type of fluids present, the
geological structure and the production mechanism. Too frequently, the time-
consuming yet intellectually mundane task of operating reservoir simulation
software overwhelms the effort to understand the field rationally: what are the
major uncertainties in the understanding of the field, what data is needed to
remove or reduce these uncertainties, what is happening now, what controls
production and, physically, what the consequences of alternative production
strategies are. The essence of good reservoir engineering is combining data,
identifying uncertainty and describing production mechanisms. It is not playing
computer games with sophisticated software as a smokescreen for a poor
understanding of the basic mechanisms by which oil is produced.
Below is a schematic of an oil field, which also contains gas, contained underneath
impermeable cap rock. The diagram is reasonable, but rather under-estimates the typical
depth of the field. Usually the oil is several km underground, while the depth of the column
of oil itself is often less than 100 m. The areal extent is generally several km2: later we will
discuss some of the world’s larger oil fields, but the total volume of oil-bearing rock is typically
around 109 m3, with, of course, a huge variation.
A schematic of an oil reservoir. The picture is reasonable, but the oil is generally found
several km below ground, while the water, oil and gas are all contained in porous rock.
The gas and oil are held in the pore spaces of the rock at high temperatures and pressures. It
is possible to estimate these values from the known depth and the geothermal gradient, as
well as the pressure gradient. A typical geothermal gradient is 30oC/km, giving temperatures
of around 100oC for reservoirs a few km deep.
The oil and gas are held in a porous rock. What does this mean? Soils, sand, gravel,
sedimentary rock and fractured rock all have some void space — that is gaps between the
solid, as shown in the figure overleaf. These systems are all porous media. If this space is
continuous, in however a tortuous a fashion, it is possible for a fluid that occupies the voids
to flow through the system – the material is said to be permeable. Soil, sand and gravel consist
of small solid particles packed together. Consolidated rock is normally found deep
underground where the individual particles have fused together. Volcanic rock that does not
naturally contain any void space can still be permeable if it has a continuous pathway of
fractures.
The fluid pressure can be estimated from the weight of fluid above it in the pore space.
pressure increases with depth as:
= + ℎ
(3.1)
where Po is a reference pressure and is the fluid density. g is the acceleration due to gravity
= 9.81 ms-2.
Putting in representative values of depth and (water) density, yields pressures of several 10s
MPa,1 or hundreds of times atmospheric pressure (which is, approximately 0.1 MPa). We will
use this equation later when it is employed to determine the depths of oil/water and gas/oil
contacts.
In modern petroleum engineering, oil fields are detected through seismic imaging, where
sound waves are sent through the rock; the returning waves detect changes in the acoustic
1
These notes will use, where possible, SI units. The US oil industry sticks doggedly to its
peculiar and nonsensical system that is often curiously described as “British”. In places I will
have to use them, in order to understand the current literature and practice. However, this is
not an excuse – often used – to employ them yourselves in serious engineering calculations
when this is not strictly necessary. Certainly the use of unit conversions in equations is
absolutely ridiculous and should never be contemplated.
properties of the rock and can be used to detect possible traps where hydrocarbons could
accumulate. It is also possible in some cases to infer directly the likely presence of
hydrocarbons.
Then an exploration well is drilled: you can never be sure that you have an oil field until you
have drilled a well and oil is produced. The seismic image may have been wrongly interpreted,
or the field may contain oil, but the flow rate is so slow as to make production uneconomic.
When the well is drilled, fluid and rock samples can be collected and brought to the surface
for further analysis.
Bentheimer (quarry)
sandstone showing
individual grains
5 mm
The first consideration is to estimate how much oil is contained in the field. This quantity is
called STOIIP (stock tank oil initially in place) and is computed as follows:
= /
(3.2)
where N is the STOIIP, is the porosity, So is the oil saturation, V is the gross rock volume and
Bo is the oil formation volume factor. Let’s go through each of the terms. The seismic image,
and the thickness of the field (or the thickness of oil-bearing rock) directly contacted by the
well, give a good inference of the extent of the field: that is the volume of porous rock that
contains oil. This is V – the gross rock volume.
However, the oil field is not an underground lake, or cavern full of oil. The oil resides in porous
rock. Only a fraction of that rock contains void space.
The porosity, , is the fraction of the volume of the porous medium occupied by void space.
This means that the porosity is the volume of void space in a soil or rock divided by the total
volume of the soil or rock (including void spaces). More strictly speaking, we mean the
effective porosity, or the volume fraction of the porous medium containing connected void
spaces through which fluids may flow. It excludes regions of void space entirely enclosed by
solid material. For most soils and unconsolidated rock the effective porosity and void fraction
are the same, but may be different for some rocks, such as carbonates and highly porous soils.
From now on when we mention porosity, we mean the effective porosity.
The porosity around 35-40% for, say, sand on a beach, but is much lower deep underground,
where the grains comprising the rock have been fused together at high temperatures and
pressures. Typical porosities lie in the range 10-25%. The porosity can be measured directly
on core samples (cm-long samples taken while drilling the well) or estimated from so-called
log or down-hole measurements.
Description Porosity %
Uniform sand, loose 46
Uniform sand, dense 34
Glacial till, very mixed-grain 20
Soft glacial clay 55
Stiff glacial clay 37
Soft very organic clay 75
Soft bentonite clay 84
The porosity of natural soils [Terzaghi and Peck, 1948] – reservoir rocks are generally
consolidated and have lower porosities typically in the range 15-30%.
Furthermore, not all the void space is full of oil. Initially, the rock is saturated with (salty)
water. Oil formed from the chemical transformation of organic matter (generally shallow
marine organisms) in deep sediments rises slowly upwards over geological time. The oil
collects under traps, from which it cannot escape, displacing the water that was initially
present. However, not all the water can be squeezed out of the rock – there is always some
water initially present. Saturation is defined as the fraction of the void space occupied by a
given phase.
The water saturation, Sw is the fraction of the void space of the soil occupied by water. The
volume of water per unit volume of soil or rock is Sw. Sw is called the moisture content
in the groundwater literature. However, here we are more concerned with systems
containing oil and natural gas.
Water saturation can, again, be measured from extracted core samples and from log
measurements of electrical resistivity (water conducts much better than gas or oil). Generally,
we see a water saturation of between 10-40% initially in oil fields.
In oil reservoirs the void spaces may contain water, oil and gas. The oil saturation is the
fraction of the void space occupied by oil and the gas saturation is the fraction of the void
space occupied by gas. The sum of the saturations of all phases is one (why?).
We now can calculate the volume of oil in the reservoir. However, when the oil is brought to
the surface – where it is sold – its volume changes. Hence, it is the universal practice in
petroleum engineering always to refer to oil volumes at so-called stock tank conditions. This
is the oil volume at a standard temperature (60oF or around 18oC) and pressure (atmospheric).
The ratio of reservoir volume to surface volume is called the oil formation volume factor. This
ratio generally lies between 1 and 2. The oil shrinks when it is brought to the surface. At first
sight, this seems counter-intuitive, since you would expect a fluid to expand as the pressure
drops. However, as the diagram overleaf explains, the oil contains dissolved gas (as we discuss
later): the exsolution of this gas as the pressure drops, means that the oil contains fewer
molecules and overall its volume decreases. Bo, the oil formation volume factor, is measured
on fluid samples taken from the well.
So, now we know how much oil we have underground – from all the terms in Eq. (3.2) – but
how do we produce it?
Oil formation
volume factor,
Bo = Vor /Vos
Pressure drops:
gas comes out of
solution
When oil flows up to the surface, its pressure drops. Bubbles of gas exsolve from the
oil. At the surface, both oil and gas are produced. The volume of oil at the surface is
lower than that in the reservoir, because gas has come out of solution. The oil formation
volume factor is the ratio of the reservoir volume of oil to the volume at standard or stock
tank conditions (60oF and atmospheric pressure).2
2
Image from dehaanservices.ca.
Oil fields are produced by drilling wells through the reservoir and allowing the oil to flow up
through the well to be collected at the surface. This process is called primary production: this
is the first process that occurs and uses the reservoir energy – essentially the pressure of the
rock and fluids – to drive out the oil. The problem with this approach to production is two-
fold. First, once the pressure has dropped sufficiently, the field will stop producing even
though it is still full of oil, so it is extremely inefficient. Secondly, and related to this, is that oil
at high pressures and temperatures is a mix of hundreds of chemical constituents and some
of these – principally the gaseous fractions methane, ethane and butane, come out of solution
when the oil is brought to the surface and the pressure drops. The pressure at which gas first
exsolves from the oil is called the bubble point. Gas has a lower viscosity than oil and so flows
much more readily; as a consequence, the gas is produced preferentially to the oil, leaving the
oil behind. Hence, in general, production is designed to maintain the reservoir pressure above
the bubble point.
How do we recover more oil, while preventing the preferential production of gas? If the well,
or wells, only produce oil, then the pressure will, inevitably, fall over time. Hence, we need to
apply so-called secondary production, where another fluid – normally gas or water – is
injected into the reservoir through injection wells. This process serves two purposes. Firstly,
it helps maintain the reservoir pressure (above the bubble point) and keeps a high driving
force to keep the oil flowing. Secondly, the water (or gas) displaces the oil from the pore
space of the rock, leading – potentially – to high recoveries.
The final process in a field life is tertiary (third) recovery, when another fluid is injected instead
of, say, just water, to remove more of the oil. Sometimes the expressions improved oil
recovery, or enhanced oil recovery are used. 3 In general, these terms refer to the injection of
something other than just water to recover as much oil as possible from the field, and are not
strictly related to the time sequence. Enhanced oil recovery can, in theory, be either a
secondary or tertiary process. Enhanced oil recovery can include the injection of gases –
natural gas or carbon dioxide – the use of foams, polymers and surfactants, and thermal
3 Generally enhanced oil recovery and improved oil recovery mean the same. However, the
word “enhanced” sounds stronger than “improved” and so is strictly reserved for the injection
of something other than normal water, while improved oil recovery is sometimes used to
cover a range of different reservoir engineering practices that the engineer considers to be
non-routine.
methods, such as steam injection, where the reservoir oil is heated, lowering its viscosity and
aiding flow. The injection of low salinity water is another example of an improved (or perhaps
even enhanced) recovery process.
Overall, a sophisticated suite of techniques can be applied to recover the oil. However, the
huge volumes of fluid involved and the oil price do severely limit the technologies that can be
employed economically. While 2008-2014 oil price of around $100/barrel was historically
high, 1 barrel is around 160 litres, so the price is only around 60 cents per litre.
As touched upon previously, around two-thirds of the oil that has been discovered is left
underground when the field is abandoned. Improvements in technology are now routinely
seeing recoveries of 50% or better, thanks to better seismic imaging of the rock (and seeing
how the images change over time), targeted drilling and the use of horizontal and slanted
wells, better simulation technologies to model the likely movement of fluids and generally a
much better understanding of flow in porous media.
This class is associated with this better understanding, laying a rigorous foundation for
improving oil recovery around the world. To help place this in perspective, the figure overleaf
shows a modern simulation model, where a structurally complex reservoir is described with
spatially varying properties (porosity and permeability, defined later). We will go through the
concepts that underlie the construction of a model of this type and the equations that are
used to solve for fluid flow. What this course is not about is how to run simulators or other
computer codes – that you can do when you are working in the industry – what you learn here
are the underlying concepts which will allow you to understand and interpret the complex and
fascinating behaviour of hydrocarbon fields.
The table overleaf (taken from Wikipedia)4 lists some of the world’s largest oil fields. While
we are still discovering oil around the world – principally deep offshore Angola, Brazil and Gulf
of Mexico – the vast majority of the world’s largest fields were discovered many decades ago.
To maintain oil production, and indeed to increase it to supply oil to a growing world’s
population with a legitimate aspiration to share in the prosperity enjoyed by the Western
world, we need to recover the oil from the fields that we have already found as efficiently as
possible; we cannot rely on finding new conventional oil. In addition, we can look for new
sources, such as shale oil (oil contained in shale or source rock), oil shale (immature source
rock where the organic material has to be heated to produce oil), or other resources. We
4 In general, I would recommend not using Wikipedia, or indeed other internet sources blindly,
but go back to the original source material. However, in some circumstances – such as this
– the information is extremely valuable (and convenient) for illustrative purposes.
currently produce around 30 billion barrels of oil each year,5 while we discover at most half
that in new fields.
The world’s largest oil field is in Saudi Arabia – Ghawar – that is undergoing the world’s largest
water injection project. It is a huge carbonate field, with fractures and zones of high
permeability (fast flow). The second largest field – Burgan – is mainly sandstone and situated
in Kuwait. Historically, the most oil of any country has been produced in the US; note,
however, that its largest oil field – Prudhoe Bay off the North coast of Alaska – is quite some
way down the list. At present, Saudi Arabia and Russia are the two largest oil producers, but
with the US catching up fast, thanks to new discoveries and shale oil.
The recoverable oil is the amount of oil that can be extracted from the field using current
technology: the amount of oil underground is typically (as I said above) some three times (or
more) larger. If you consider that the oil price is $100/barrel, then a 1 billion barrel oil field
represents $100 billion of potential value. Imagine that a sensible application of the ideas in
this class could improve overall recovery of, say, 1% for a single field; this represents a
considerable amount of money!
5 A great source of public domain information on oil (and other energy) production is the BP
statistical review of energy - http://www.bp.com/en/global/corporate/about-bp/energy-
economics/statistical-review-of-world-energy-2013.html .
Production
Started Recoverable Oil
Field Location Discovered (Million
Production (Billion Barrels)
Barrels/day)
Ghawar Field Saudi Arabia 1948] 1951 75-83 5
Burgan Field Kuwait 1937 1948 66-72 1.7
Ferdows/Mound/ 7-9 (38 Gb
Iran 2003
Zagheh Field resource)
Sugar Loaf field Brazil 2007 possibly 25-40
Cantarell Field Mexico 1976 1981 18 .408
Bolivar Coastal
Venezuela 1917 1922 30-32 2.6-3
Field
Azadegan field Iran 2004 9
Brazil, Santos
Lula Field 2007 5-8
Basin
Saudi
Safaniya-Khafji
Arabia/Neutral 1951 30
Field
Zone
Esfandiar Field Iran 30
Rumaila Field Iraq 1953 17 1.3
Tengiz Field Kazakhstan 1979 1993 26-40 .53
Ahwaz Field Iran 1958 10.1 .700
Kirkuk Field Iraq 1927 1934 8.5 0.480
Shaybah Field Saudi Arabia 15
Agha Jari Field Iran 1937 8.7 .200
Majnoon Field Iraq 1975 11-20 0.5
Russia, West
Samotlor Field 1965 1969 14-16 0.844
Siberia
Romashkino Russia Volga-
1948 1949 16-17 .301 (2006)
Field Ural
United States,
Prudhoe Bay 1969 13 0.9
Alaska
12 (6.5 billion
Sarir Field Libya 1961 1961
recoverable)
Russia, West
Priobskoye field 1982 2000 13 0.680 (2008)
Siberia
Lyantorskoye Russia, West
1966 1979 13 0.168 (2004)
field Siberia
Abqaiq Field Saudi Arabia 12 0.43
Chicontepec
Mexico 1926 6.5 (19 certified)
Field
Berri Field Saudi Arabia 12
Zakum Field Abu Dhabi, UAE 1965[25] 1967 12
0.18-0.25
West Qurna Field Iraq 1973 15-21
(potential)
Manifa Field Saudi Arabia 11
Fyodorovskoye Russia, West
1971 1974 11 1.9
Field Siberia
East Baghdad
Iraq 1976 8 0-0.05 (potential)
Field
Faroozan-Marjan
Saudi Arabia/Iran 10
Field
Brazil, Campos
Marlim Field 10-14
Basin
Awali Bahrain 1
Aghajari Field Iran 14
Azadegan Field Iran 1999 5.2
Gachsaran Field Iran 1927 15
Marun Field Iran 16
Mesopotamian
Kuwait 66-72
Foredeep Basin
Production
Started Recoverable Oil
Field Location Discovered (Million
Production (Billion Barrels)
Barrels/day)
Minagish Kuwait 2
Raudhatain Kuwait 11
Sabriya Kuwait 3.8-4
Yibal Oman 1
Dukhan Field Qatar 2.2
Halfaya Field Iraq 4.1
Az Zubayr Field Iraq 6
Nahr Umr Field Iraq 6
Abu-Sa'fah field Saudi Arabia 6.1
Hassi Messaoud Algeria 9
Kizomba
Angola 2
Complex
Dalia (oil field) Angola 1
Belayim Angola >1
Zafiro Angola 1
Zelten oil field Libya 2.5
Agbami Field Nigeria 0.8-1.2
Bonga Field Nigeria 1.4
Azeri-Chirag-
Azerbaijan 1985 1997 5.4
Guneshli
Karachaganak
Kazakhstan 1972 2.5
Field
Kashagan Field Kazakhstan 2000 30
Kurmangazy
Kazakhstan 6-7
Field
Darkhan Field Kazakhstan 9.5
Zhanazhol Field Kazakhstan 3
Uzen Field Kazakhstan 7
Kalamkas Field Kazakhstan 3.2
Zhetybay Field Kazakhstan 2.1
Nursultan Field Kazakhstan 4.5
Ekofisk oil field Norway 3.3
Troll Vest Norway 1.4
Statfjord Norway 3.4
Gullfaks Norway 2.1
Oseberg Norway 1979 1988 2.2 3.78
Snorre Norway 1.5
Mamontovskoye
Russia 8
Field
Russkoye Field Russia 2.5
Kamennoe Field Russia 1.9
Vankor Field Russia 1983 2009 3.8
Vatyeganskoye
Russia 1.4
Field
Tevlinsko-
Russkinskoye Russia 1.3
Field
Sutorminskoye
Russia 1.3
Field
Urengoy group Russia 1
Ust-Balykskoe
Russia >1
Field
Tuymazinskoe
Russia 3
Field]
Arlanskoye Field Russia >2
South-Hilchuy
Russia 3.1
Field
Production
Started Recoverable Oil
Field Location Discovered (Million
Production (Billion Barrels)
Barrels/day)
North-
Russia 2.2
Dolginskoye Field
Nizhne-
Russia 1.7
Chutinskoe Field
South-
Russia 1.6
Dolginskoye Field
Prirazlomnoye
Russia 1.4
Field
West-
Matveevskoye Russia 1.1
Field
Sakhalin Islands Russia 14
Odoptu Russia 1
Arukutun-Dagi Russia 1
Piltun-
Astokhskoye Russia 1
Field
Ayash Field East-
Russia 4.5
Odoptu Field
Verhne-
Russia 1.3
Chonskoye Field
Talakan Field Russia 1.3
North-Caucasus
Russia 1.7
Basin
Clair oil field United Kingdom 1977 1.75
Forties oil field United Kingdom 1970 5
Jupiter field Brazil 7
Cupiagua/Cusian
Colombia 1
a
Boscán Field,
Venezuela 1.6
Venezuela
Pembina Canada 1953 1953
Swan Hills Canada
Rainbow Lake Canada
Hibernia Canada 1979 1997 3
Terra Nova Field Canada 1984 2002 1.0
Kelly-Snyder / United States,
1.5
SACROC Texas
3.0 (2.0 billion
United States, recovered; 1.0
Yates Oil Field 1926 1926
Texas reserve
remaining)
United States,
Kuparuk oil field 1969 6
Alaska
United States,
Alpine 0.4-1
Alaska
East Texas Oil United States,
1930 6
Field Texas
United States,
Spraberry Trend 1943 10
Texas
Wilmington Oil United States,
1932 3
Field California
South Belridge United States,
1911 2
Oil Field California
United States,
Coalinga Oil Field 1887 1
California
United States,
Elk Hills 1911 1.5
California
United States,
Kern River 1899 2.5
California
Production
Started Recoverable Oil
Field Location Discovered (Million
Production (Billion Barrels)
Barrels/day)
Midway-Sunset United States,
1894 3.4
Field California
Thunder Horse United States,
>1
Oil Field Gulf of Mexico
Kingfish Australia 1.2
Halibut Australia 1
Daqing Field China 1959 1960 16
Fluid mass above a point increases with depth and thus pressure increases with depth as given
by Eq. (3.1). This can be written in terms of a pressure gradient with z referring to depth:
(3.3)
Pressure increases fastest for water with depth, then oil, then gas.
In oil field operations the fluid pressure can be measured downhole using an RFT (repeat
formation tester). Consider the schematic diagram of pressure as a function of depth overleaf.
Pressure measurements can be used to identify the locations of the contacts between oil and
water, and gas and oil, even if these are not detected directly by logs.
If the pressure profiles from different wells do not overly as a function of depth, then this can
imply the presence of non-communicating regions (shales and faults) indicating reservoir
compartmentalization.
As the diagram indicates, the pressure in the hydrocarbon-bearing zones is higher than in the
surrounding aquifer (water-saturated rock). This has nothing to do with compressibility, but
is caused simply by the density differences between the fluids. The water and oil have the
same pressure at the free water level: the oil pressure decreases less rapidly with height
(decreasing depth) than water, and so the oil pressure is higher than the water pressure above
the free water level. The same is true for the gas column. There is an analogy with keeping a
balloon underwater in the bath – you need to force the balloon down to prevent it rising as it
is buoyant. This is equivalent to the fluid pressure in the cap rock necessary to stop the oil
and gas escaping upwards.
Depth, z
Pressure difference
GAS
Free oil level
OIL
Pressure
A schematic of the pressure profile with depth for an oil field with a gas cap. The oil and
water pressures are equal at the free water level. The pressure difference – where the
gas pressure is higher than that in the surrounding aquifer – is indicated at the top of
the reservoir.
The higher pressure in the hydrocarbon-bearing zone compared to a water column at the
same depth poses a potential hazard during drilling. A water-filled drill hole, when it
encountered oil or gas, would allow a blow-out: the higher pressure hydrocarbon would flow
into the hole and rise uncontrolled to the surface. This is the principal reason for the use of
drilling mud – dense fluid mixed with water or oil – which ensures a higher pressure in the
well than the formation. The mud also cools and lubricates the drill bit and helps carry cuttings
to the surface.
Now we consider the same diagram, but where specific pressure measurements have been
made at the depths indicated. If we know the densities of the fluids (these can be obtained
from measurements on downhole samples, or – more simply – from the slopes of measured
pressure as a function of depth – the depths of the free water and oil levels can be computed.
Depth, z
Pressure difference
z1 GAS
Free oil level, zo
z2 OIL
Pressure P1 P2 P3
A schematic of the pressure profile with depth for an oil field with a gas cap where
specific pressure measurements at the indicated depths are shown. From this the
depths of the free water and oil levels can be computed.
Then using the measurements together with Eqs. (3.1) and (3.3) we can write for the gas, oil
and water pressures respectively:
= +( − )
(3.4)
= +( − )
(3.5)
= +( − )
(3.6)
The free water level is – by definition – when the oil and water pressures are the same. So,
equating Eqs. (3.5) and (3.6) we find:6
6 One tip for the calculation: get your units straight. If depth is in m and density in kgm-3, then
you need to do the calculation in Pa (not MPa or anything else, such as – horrors – psi, a
senseless unit which should be forever banned from petroleum engineering.)
− + −
=
( − )
(3.7)
Similarly, for the free oil level where the oil and gas pressures are the same:
− + −
=
( − )
(3.8)
The height of the oil column – used to determine the initial oil in place – is simply zw-zo.
The other way to determine the contacts between phases and the height of the oil column, is
directly from log measurements. Here downhole readings of resistivity (to distinguish
between water and hydrocarbons) and density (to distinguish between oil and gas) can be
used to locate the depths of the contacts. In practice, however, the well may not penetrate
the contact directly, or the readings may be ambiguous or open to different interpretations.
The essence of reservoir engineering is to consider all the data together – there is never one
overriding DNA-type test that trumps all others – so you must always assess all the evidence
carefully and arrive at a determination that is consistent with it.
There is a difference anyway between the free water level determined from the pressure and
the contact determined from logs – this difference is the capillary pressure. In a reservoir that
is initially water-wet, a pressure difference between oil and water (that is, a higher pressure
in the oil) is necessary for oil to enter the porous medium during primary oil migration. This
means that the true contact – where the oil saturation is significant – generally lies above the
point where the oil and water pressures are the same – where the capillary pressure (the
difference between the oil and water pressures) is equal to the capillary entry pressure. In
general – for permeable sandstones – this difference is less than 1 m, but it may be significant
in lower permeability systems. A full understanding of this requires a knowledge of capillary
pressure, provided in the following sections.
A normally pressured reservoir. The water pressure is as expected for its depth – there is a
continuous pathway of water in the pore space to the surface. If z is the depth from either
the water table (onshore) or the sea surface (offshore) then the water pressure is given by
= + , where Patm is atmospheric pressure and we have assumed a constant
water (brine) density with depth. Note that we only consider the water pressure – the oil and
gas pressures are higher than water for a given depth, as discussed above.
Over-pressured. The reservoir pressure higher than expected – that is the real water pressure
is higher than you would expect for a normally pressured reservoir. This means that the
reservoir has been uplifted since filling. Note that this is the pressure relative to expected for
depth – it is not an absolute measure. In an over-pressured reservoir, production may be
rapid, but during drilling there is a risk of blow-outs, as the reservoir pressure could be higher
than the mud pressure downhole.
Under-pressured. This is when the water pressure lower than expected. The reservoir has
been down thrown over geological time: more sediment has been deposited over the field
since it was charged with oil and gas.
We have already defined the oil formation volume factor. In this section we define some
other terms necessary for our analysis of oil production. We extend the diagram we showed
previously, to include the more general case where both oil and gas are produced from the
reservoir – that is, oil and gas are both present in the reservoir itself.
The various terms are easy to define using equations based on the figure above, but require
great care when described in words.
So, for completeness, here are the definitions. The oil formation volume factor Bo, is the ratio
of the reservoir volume of oil to the surface volume of oil:
(3.9)
Traditionally oil volumes are measured in barrels. To emphasise the nature of the conversion,
oil volumes measured at the surface are called stock tank barrels (stb), while volumes in the
reservoir are measured in reservoir barrels (rb). Hence the common unit for Bo – while it is
strictly dimensionless – is rb/stb. A typical range is from 1 (heavy oils with little or no
associated gas) to more than 2 for highly volatile oils.
Gas at the
surface,
Surface Oil at the surface, volume Vgs
volume Vos
Solution + Free gas
Oil formation gas Vsg Vfg
volume factor,
Bo = Vor /Vos
Pressure drops:
gas comes out of
solution
A schematic showing the production of oil and gas from a reservoir with the
corresponding volumes of oil and gas at the surface. The oil volume shrinks because
of the exsolution of dissolved gas: the gas produced is both solution gas (from the oil)
and free gas (which was gas in the reservoir as well).
A saturated oil is defined as one that cannot hold any more gas. This means oil at, or below,
the bubble point, when gas first comes out of solution. The volume of the oil decreases as the
pressure drops.
An undersaturated oil is one that can dissolve more gas if it is available. The volume of oil
increases as the pressure drops, until the bubble point is reached (the oil becomes saturated).
The terms saturated and undersaturated are commonly used, but I find them confusing. It
shows more clarity and confidence to refer simply to an oil that is above or below the bubble
point.
The gas formation volume factor, Bg, is the reservoir volume of free gas divided by the surface
volume of that free gas (the solution gas is not included):
(3.10)
Again the units here should be dimensionless and indeed they are in SI units (rm 3/sm3 where
the r and s stand for reservoir and surface respectively), in field units the situation is a little
more confused. Reservoir volumes of both oil and gas are measured in rb, while gas volumes
at the surface are measured in cubic feet (scf). So the units of Bg are traditionally rb/scf.7
R is the producing gas/oil ratio, or the ratio of the total gas production rate to the oil
production rate, with both volumes measured at surface conditions:
+
= =
(3.11)
The field units for R are scf/stb. Note that strictly this is a ratio of rates – or volumes produced
over a fixed, short, time period, such as one day. It is not a ratio of total volumes since the
field was produced – this is defined and used later.
Rs is the solution gas/oil ratio and is the volume of gas (measured at standard conditions) that
will dissolve in unit surface volume of oil in the reservoir:
(3.12)
For gases, we often do not quote the formation volume factor directly, but invoke a gas law.
The ideal gas law – with which I hope you are familiar – is:
=
(3.13)
where P is pressure, V is (gas) volume, n is the number of moles and T is the absolute
temperature (Kelvins – I don’t even know what silly absolute temperature unit has been
7 At this point you may think that some conversion factors may be helpful. Some are provided
in the Appendix, but when I need them I simply look them up on the internet – this is one
occasion where this is appropriate and convenient.
invented for the Fahrenheit scale). R in this context is the universal gas constant whose value
(in SI units it is 8.314 JK-1mol-1 – here I simply refuse to contemplate its value in inconsistent
should-have-died-out-with-the-dinosaurs-units).
The ideal gas law assumes that the molecules in the gas behave as point particles with no
mutual interaction. This is a good approximation at atmospheric temperatures and pressures
(stock tank conditions) but the nature of the “gas” encountered in the reservoir is very
different. At pressures that are hundreds of times atmospheric the gas has a density
comparable to oil and there is significant interaction between the gas molecules. Indeed, the
distinction between gas and liquid is not obvious – strictly we refer to gas as the less dense
phase and oil as the more dense phase when two hydrocarbon phases are present in the
reservoir.
=
(3.14)
It is possible to determine Bg from Z using Eq. (3.14). If we assume that Z=1 at surface
conditions (denoted by the subscript s) then, since when the gas expands from reservoir to
surface conditions the number of moles is the same:
(3.15)
= =
(3.16)
The figure overleaf shows typical values of Bg, Bo and Rs as a function of pressure for a typical
North Sea oil.
Note that Bo reaches is maximum value at the bubble point. Above this pressure, oil expands
as the pressure drops (like a normal fluid) and so Bo rises as pressure drops. Below this
pressure, however, the exsolution of gas causes the oil to shrink as the pressure drops, and Bo
correspondingly falls.
Values of the oil formation volume factor, the gas formation volume factor and the
solution gas/oil ratio (called gas solubility) for a typical North Sea oil.8 The oil formation
factor is maximum at the bubble point – this is also the pressure below which the gas
solubility falls.
The range of Rs in hydrocarbon fields is considerable, from infinite (for a dry gas – as gas that
produces no liquid at the surface) to zero (a heavy oil with no associated gas). It is useful to
have some concept of whether or not a field is producing more oil or gas. From a volume
perspective, the gas wins, since it is much less dense than oil at the surface. This is further
confused by the baffling and unhelpful units of scf/stb (or MScf9/stb). So let’s consider this on
a mass basis. Imagine that the gas is largely methane, then we can estimate the density at
standard conditions from the ideal gas law, Eq. (3.13). The molecular mass, M, of methane is
0.016 kgmol-1 (note the units – don’t try some sort of idiotic compromise over units and
muddle yourself with a molecular “weight” in g/mol – this won’t work and you know it won’t
work) and so the density is given by:
= =
(3.17)
We now compare this with typical oil densities (at surface conditions). I will take a
representative value of 800 kgm-3.10
If a hydrocarbon field produces the same mass of gas as oil, then what is the value of Rs? In
SI units this would be 800/0.67 = 1,200 m3/m3. In field units, since 1 m3 = 6.2898 stb and 1 ft
= 0.3048 m, we have a ratio of 1.2/(6.2898 × 0.30483 ) =1 stb = 6.7 Mscf/stb.
So, a field with an Rs of greater than around 6-7 Mscf/stb is gas-like, while one with a smaller
gas/oil ratio is oil-like. In the next section we provide strict thermodynamic definitions, but
this is a useful guide. Wet gases (that produce some liquid at the surface) generally have Rs
around 50 Mscf/stb or larger; gas condensates are in the intermediate range 5-30 Mscf/stb;
volatile oils 2-3 Mscf/stb and black oils (we define these terms later) 0.1-2 Mscf/stb. I will
9
Another stupidity with field units. M stands for million (106) of course – or at least in SI units.
In field units M stands for the French word “mille” or thousand – hardly British units? Million
is represented by MM. Confusing – stick to the M and MM convention for field units, but don’t
get into a silly muddle and do the same when SI units are (grudgingly) applied.
10
In petroleum engineering it is common to categorize an oil by its density. Fine, since this
does tend to correlate with composition and viscosity as well. However, the unit is bizarre
and essentially unusable in any engineering calculation: API gravity (API stands for the
American Petroleum Institute, an organisation, in my opinion, of dubious utility). Needless to
say, if you have got this far in the notes and want to use API gravity, it is time to move to
another text book. API gravity will not feature again here.
repeat that these are simply rough estimates and should never be used to define a producing
field.
The diagram on the next page provides a strict thermodynamic definition of the different
types of hydrocarbon that are produced from conventional oil and gas fields. The diagram
shows the phases present for a given hydrocarbon composition as a function of pressure (the
y axis) and temperature (x axis). Within the region indicated two hydrocarbon phases (oil and
gas) are present; outside this region only one phase is present.
This phase diagram is not known a priori; it needs to be computed based on experimental
measurements on fluid samples taken from the reservoir. Normally, measurements are used
to tune an equation of state model, from which this phase diagram can be determined; further
discussion lies outside the scope of these notes.
Now consider a field that has just been discovered – it has a high pressure initially. What
happens as the field is produced and the pressure drops? To a very good approximation we
assume that the reservoir temperature stays constant (it can only change the temperature if
we inject a considerable quantity of liquid that is hotter, say steam, or cooler, say water, than
the reservoir, or induce thermal reaction, such as in situ combustion). In a gas field – in the
reservoir itself – the fluid expands without a phase transition. This is the definition of a gas
field – no phase change in the reservoir as the pressure drops. Now, at the surface – at a
lower temperature – some liquid (oil) may be produced (it condenses from the gas stream).
A gas field that produces liquid oil at the surface is called a wet gas (note that this has
absolutely nothing to do with water production), while a gas field that produces no liquid is
called a dry gas.
A gas condensate field produces oil in the reservoir itself as the pressure drops. How can
we define this as a gas field, as opposed to an oil field? Not on some ad hoc limit on the value
of Rs. No, because when there is a phase transition, the first drop of the second phase to
appear is denser than the majority phase: we define this denser phase as oil and the less dense
phase as gas. The pressure at which this occurs is called the dew point. Remember that we
are dealing with high pressures, and the densities of the oil and gas may be similar.
An oil field produces gas as the pressure is dropped in the reservoir: we know that this is gas,
as the first bubbles of the second phase are less dense than the original, oil, phase. The
pressure when this occurs is the bubble point.
Two-phases present
A schematic phase diagram for a hydrocarbon mixture for different pressures and
temperatures. The mixture separates into two phases – oil and gas – in the region
indicated. The red spots mark the initial reservoir conditions for different types of
hydrocarbon field; the green spots label the surface conditions.
The critical point marks the transition from an oil field to a gas condensate and is the point
where the oil and gas properties are indistinguishable. To the left of the critical point on the
diagram, we have an oil field; to the right we have a gas condensate field.
The fluid properties have an impact on production mechanisms. In a gas field, the preferred
method of production is simply to drop the pressure. The gas expands in the field with no
phase separation in the subsurface: if you drop the pressure to close to atmospheric,
recoveries of 90-95% are possible (see the section on Material Balance later). In contrast, in
an oil field, when the pressure drops below the bubble point, gas is formed. This gas is less
viscous than the surrounding oil and – once it is connected in the pore space – will flow
preferentially to the oil. Thus, you produce gas and leave behind the (more valuable) oil. Even
if you drop the pressure to near atmospheric, you still have a field full of oil and, since the oil
is not very compressible, this results in rather poor recovery factors (typically around 20%).
So, for an oil field, it is necessary to maintain the pressure above the bubble point, normally
through the injection of water.
A gas condensate is more complex. The lowest pressures are near the well. If the pressure
drops below the dew point, oil is formed. Once again the more valuable fraction of the
hydrocarbon – the liquids – are left behind in the reservoir. Moreover, the liquid blocks the
pore space and reduces the productivity of the well. There are two options. One option is to
maintain pressure, as in an oil field, or produce more slowly, although this has an economic
cost. The other is to drop the pressure further until we are back in the single-phase (gas)
region – see the previous diagram. Then the reservoir can be produced as a normal gas field.
The phase diagram is often used to describe different types of oil field. A black oil is not
specifically a reference to the colour of crude oil (which indeed is black) but to how it may be
described thermodynamically. Most fields are characterised as having an oil with dissolved
gas with a solution gas/oil ratio Rs defined as a function of temperature and pressure. The gas
and oil also have properties that vary only with temperature and pressure. This seems
obvious, but in reality, the exsolved gas has a different composition, dependent on
temperature and pressure and production process – after all the overall hydrocarbon
composition of the reservoir changes if gas is, for instance, preferentially produced compared
to gas. For most fields it is a reasonable approximation to assume that the gas has a fixed
composition, independent of production path. However, for fields with a temperature near
the critical temperature, this is not so accurate. These can be described a volatile or near-
critical oil fields and often a more sophisticated compositional characterization of their
behaviour is required. Furthermore, for gas injection, we need also to account for gases that
have distinct properties – the injected gas will not normally have a composition identical to
exsolved gas. In contrast, a heavy oil may have little or no dissolved gas present.
The diagram we have presented assumes a fixed hydrocarbon composition in the reservoir
and explores the behaviour for different initial temperatures. In reality, the variation in initial
reservoir temperature is less dramatic than changes in composition – which can vary from
pure methane (a dry gas) to virtual solids (oils with a composition dominated by molecules
with chains of C30 – that is, 30 carbon atoms – or longer).
The following diagram assumes that the initial and surface conditions are fixed and shows the
effect of composition on the phase diagram. The figure should be self-explanatory and is
consistent with the definitions provided previously. Note that the phase diagrams tend to
shift downwards (two-phase conditions at lower pressures) and to the right (two-phase
conditions at higher temperatures) as the composition varies. Typically, a dry gas is almost
entirely methane (C1), wet gas contains significant quantities of C1-C6, gas condensates and
near-critical oils contain a range of hydrocarbons (say C1-C10) with large quantities of C1-C4 and
some heavier components, for a black oil C1-C20 is most common, while heavy oils are
predominately C10 and heavier.
Pressure
Initial conditions
Dry gas
Wet gas
Gas condensate
Near-critical oil
Black oil
Heavy oil
Surface conditions
Temperature
Schematic phase diagrams for different types of hydrocarbon reservoir. The red spot
marks the initial reservoir conditions; the green spot labels surface conditions. The
black spots mark the critical points for the different types of field.
This concludes a simple overview of hydrocarbon phase behaviour – for further reading the
classic work in this area is The properties of petroleum fluids by W D McCain, PenWell Books,
second edition (1990). What we have described is sufficient though for the discussion of
material balance in the next section.
4. MATERIAL BALANCE
As an introduction, consider the diagram below, giving a schematic overview of the changes
that occur during production.
Production wells
Pressure, Pi, and saturation, Soi
Pressure, P, and saturation, So
Schematic diagram showing an oil field at initial pressure and saturation (left), and at
sometime later (right), during production, when the oil saturation and pressure have
decreased. The bold vertical lines represent production wells.
In this example, the gross rock volume is V and the average porosity is . Then the initial
reservoir volume of oil is VNGSoi. The subscript “i” refers to initial conditions. We have added
a new concept here – NG or the net-to-gross. Traditionally, the porosity and oil saturation are
only defined for the portions of the gross rock volume (a volume that encompasses all the oil
present in the reservoir) that contain producible hydrocarbon (that is oil that can reasonably
flow to a well). The fraction of the gross rock volume that has sufficient porosity, oil saturation
and permeability to allow significant production is called the net-to-gross. What exactly is
meant by “sufficient” porosity etc? This is normally empirically defined by petrophysicisits
using log analysis and experience, and is open to interpretation. However, a discussion of this
topic lies outside the scope of these notes.
= /
(4.1)
During production the pressure drops as oil is produced. Moreover, the oil saturation may
decrease, as gas comes out of solution, and/or water is injected or ingresses from a connected
aquifer. The remaining volume of oil – measured at stock tank conditions – sometime later is
given by:
− = /
(4.2)
In this equation Np is the cumulative oil produced (measured at surface conditions, of course).
Hence, the oil volume remaining in the reservoir is N-Np, as indicated in Eq. (4.2).
The recovery factor, Rf, is defined as the ratio of the oil produced to the initial oil present:
= /
(4.3)
1− =
(4.4)
=1−
(4.5)
In any recovery process we wish to make the recovery factor as high as possible, subject to
economic and engineering constraints. How do we do this? From Eq. (4.5) we wish to make
the oil saturation So as low as possible – clearly. We need to remove as much oil as possible
from the pore space. In general, the oil saturation is a function of what and how other fluids
are injected or ingress into the reservoir, and is largely independent of pressure. But what
about pressure? From Eq. (4.5) it is also evident that we wish Bo to be as large as possible. If
we had a gas field, this simply means that we drop the pressure as low as feasible and allow
the gas to expand. But, for an oil field, the situation is different. As explained in the previous
section, Bo is a maximum at the bubble point: regardless of the injection strategy therefore, it
makes sense to drop the pressure to, but not below, the bubble point pressure.
This was a first example to show how a simple accounting for volume, combined with
conversion from reservoir to surface, can be used to obtain useful, quantitative insights into
reservoir management. We will now apply more sophisticated versions of the material
balance equation for both oil and gas fields.
It is unusual for a reservoir simulation study to be performed for gas fields – particularly
smaller fields. Hence, material balance remains the principal method for engineering analysis
in this case.
Assume that we have a dry gas – no oil produced, and no aquifer movement. If there is no
aquifer movement, then we can assume that the gas saturation remains constant; we will
relax this approximation later. We repeat the diagram shown previously, but simply change
the nomenclature for gas (we replace the subscript “o” by “g”).
Schematic diagram showing a gas field at initial pressure and saturation (left), and at
sometime later (right), during production, when the pressure has decreased. We assume
no aquifer movement in this case and so the gas saturation remains the same.
If you can immediately derive the material balance equation, similar functionally to Eq. (4.5),
then fine. However, it is always instructive to go through this step by step until the derivation
comes easily. If you know how to derive these equations, then you do not need to learn by
rote what the equations are for every specific case, and you have the ability to derive new
equations for new circumstances (shale gas, gas condensates?).
The initial gas in place measured at surface conditions, G, is given by (see Eq. (4.1)):
= /
(4.6)
The remaining volume of gas – measured at stock tank conditions – sometime later is given
by:
− = /
(4.7)
Gp is the cumulative oil produced (measured at surface conditions, of course). Therefore, the
gas volume remaining in the reservoir is G-Gp, Eq. (4.7).
In this case the recovery factor, Rf, is defined as the ratio of the gas produced to the initial gas
present: = / . Then from Eqs. (4.6) and (4.7) we obtain:
= = 1−
(4.8)
Traditionally, this equation is written in terms of the Z-factor. Using Eq. (3.16) we find:
= =1−
(4.9)
For a producing field, we wish to know how much gas was originally in place (G) and how much
gas will be produced at a certain pressure (Gp).
The way in which this is done is through the “P/Z” plot – a universal analysis for gas fields:
plot Gp vs. P / Z. From Eq. (4.9), the slope is GZi / Pi and intercept at P = 0 is G.
Let’s consider a simple example: Zi = 0.8 and Pi = 200 atm. The remaining production data is
shown in the table below.
Gp P Z P/Z
(108 scf) (atm) (atm)
0 200 0.80 250
1.53 180 0.85 212
2.56 160 0.86 186
3.78 140 0.90 156
Let’s consider a problem where we find G, and Gp at the abandonment pressure of 50 atm
when Z = 0.95.
Gp (108 scf)
At P=50 atm, Z=0.95,
P/Z=52.6 atm
0
0 50 P/Z (atm) 250
Classic P/Z plot for the analysis of a gas field using material balance, for the simple
example described in the text.
The mechanics of this example are simple. This is a useful analysis, as it allows gas sales
contracts to be made, based on a sound production-based estimate of reservoir size. The
problem is that you do need some significant period of production before reliable estimates
can be made.
Production data. Gp is the cumulative gas produced from one connected reservoir, not a
single well. Often P/Z plots on a well-by-well basis are studied, assuming that to a good
approximation each well drains a separate section of the field. However, this is not how the
material balance equation was developed. We discuss later how material balance can be used
to determine compartmentalization of the field.
Pressure. Pressure can be measured down-hole during production. However, while the well
is producing, the pressure is lowest at the well. Strictly, the pressure used in the material
balance equation is the average for the entire field: this is the pressure at the well when there
is no flow, at equilibrium. The average reservoir pressure is normally determined during
periods of shut-in using standard procedures in well test analysis (that is, using a flow model
to extrapolate the pressure response to infinite time).
The abandonment pressure is not 1 atm. Why is this? Additional pressure is required for the
gas to flow to the surface, through the surface equipment and along pipelines. This pressure
is also related to the desired (or contracted) production rate. Gas can be compressed and
pumped, but this requires additional costs and energy, which have to be compared with the
benefits of additional production. Even so, recovery factors can be very high – 79% - in this
case, and higher in many other fields, as the gas expansion can be considerable.
Z factor. As mentioned previously, the Z factor can be measured directly in the laboratory
from gas samples taken from the well. More usually, it is estimated using correlations tuned
to data and the measured composition of the gas. It needs to be known as a function of
pressure at the reservoir temperature. If data is unavailable, it is possible to estimate Z (or
Bg) from a sensible extrapolation from previous values.
Consistency with other data. As mentioned at the outset, the essence of good reservoir
engineering is to assess information from different sources. If there is a seismic survey, then
– in principle – G can be estimated from this survey, a determination (usually from logs) of the
gas/water contact and log measurements of porosity, saturation and net-to-gross. This needs
to be compared with the value obtained from material balance. If the two values are
consistent within the bounds of uncertainty, this gives confidence that the production wells
are contacting all the field. If the seismic value is considerably greater, this implies that the
field could be compartmentalized – that is the production wells are not contacting all the
producible gas and that the field is divided into distinct compartments or reservoirs. If the
seismically-derived value is smaller, then – assuming that the seismic data is sound – then this
implies that there are other mechanisms contributing to production (such as aquifer influx)
which need to be considered. This is addressed next.
Our analysis has ignored any contribution to recovery from the expansion of connate (initial)
water in the pore space and the compaction of the rock itself. Compressibility measures the
fractional change in volume per unit decrease in pressure. Technically, this is called
“compressibility,” even though in our case the fluids are expanding. An analogy would be the
air in a bicycle tyre. If you release the pressure, the air expands and flows out of the valve.
More compressible fluids expand more as the pressure is dropped, pushing out more oil or
gas. In the reservoir, the fluid pressure drops and the fluids (oil, gas and water) expand,
pushing oil and gas out through the rock pores and into the well. The rock also gets
compressed – like squeezing a sponge to release water – and this adds to the production.
Unlike sponges or air, the compressibility of the rock and fluids we consider is much lower, so
the change in volume is relatively smaller.
The figure overleaf shows a schematic of what is meant by the compressibility of fluids and
rock, and how it contributes to production. As the pressure drops, the hydrocarbon and water
expand. The rock grains – held apart by the high fluid pressure – begin to crush together,
resulting in a decrease in porosity.
We have explicitly accounted for gas (and oil) compressibility through the use of formation
volume factors, which convert reservoir to surface volumes at different pressures, but we also
need to account for the additional production from the expansion of water and the decrease
in porosity.
The pore volume of gas, Vg = (1-Swc)V, where Swc is the connate water saturation (strictly the
average initial saturation in the reservoir and Sg = 1- Swc). Vg can be written as: Vg = Vp – Vw,
where Vp is the pore volume V and Vw is the water volume SwcV. The change in pore volume
with pressure has two contributions from the compressibility of the formation and water
expansion:
Initial pressure
Pressure decreases
Fused sand grains Water expands Decrease in porosity that
press closer together squeezes out more
Hydrocarbon
hydrocarbon
expands
expands
Current pressure
= − = −
(4.10)
The first term on the right-hand-side of Eq. (4.10) represents the rock compressibility and the
second the expansion of water.
1
=−
(4.11)
The minus sign indicates that in normal substances, volume decreases as pressure increases.
1
=−
(4.12)
1
=
(4.13)
Note that there is no minus sigh, as porosity decreases as pressure decreases. More precisely,
this is the pore volume compressibility induced by a decrease in pore pressure, with
compaction only in the vertical direction. For further discussion of this point, I recommend
Compressibility of Sandstones by R W Zimmerman, Elsevier Science Publishers, New York, NY,
USA, ISBN: 0444-88325-8 (1991).
Rock compressibilities vary from values similar to – or even lower than – water for highly
consolidated sandstones, to values as high as 10-7 Pa-1 for unconsolidated sands.
We now return to Eq. (4.10) which can be written in terms of the compressibilities of water
and rock as:
1 +
=
1−
(4.14)
The fractional change in pore volume as a result of the two effects of rock compaction and
water expansion are usually small and largely outweighed by the gas expansion.
Consider, for instance, an example where Pi = 300 atm and Swc = 0.2 with a pressure drop of
100 atm (P = 107 Pa). If we take cw = 5×10-10 Pa-1 and c = 10-9 Pa-1 and assume that they are
constant as a function of pressure, then from Eq. (4.14) the fractional change in volume can
be written:
Δ +
= ΔP~ 1.4%
1−
(4.15)
Rock compressibility contributes about 1% to the change in volume and initial water
expansion 0.1%: these are relatively small effects.
Another way to see this is to consider the non-ideal gas equation, (3.14). From this and the
definition of compressibility, Eq. (4.11) we can write:
1 1 1
=− = +
(4.16)
A useful guide is that – taking the first term in Eq. (4.16) – a gas compressibility is
approximately given by the inverse of pressure. So, for a representative pressure of 200 atm
(2×107 Pa) the compressibility is 5×10-8 Pa-1, which is 50 times greater than rock and 100 times
greater than water.
We will now ignore the (usually small) effects of rock and connate water compressibility and
consider the impact of aquifer influx on the behaviour.
Before the equations are presented, it is valuable to emphasise the approach that is taken in
this course. Firstly, why are we now ignoring rock compressibility? Surely it is more “accurate”
to include it? Reservoir engineers, often goaded by elaborate software, often consider than
a model with more variables and parameters is superior to one that is simpler. However,
often the story of the field – the production mechanism and the controls on recovery – can be
obscured behind a number of poorly-understood inputs. You can ignore, for instance, rock
compressibility if it is a small effect, or include it, but you should in both cases have a clear
understanding of its impact. Secondly, we will include it in the full material balance equation
presented later. In any event, it is preferable to understand, clearly, the concepts than simply
and unthinkingly apply complex equations or software.
In this portion of the notes, we consider aquifer influx. We assume that connected to the gas
field is a large body of water contained in porous rock. As the gas pressure decreases, water
moves into the gas field in response. We will consider only one, simple, model of this
response: there are many other aquifer models that can be used, and I recommend the
textbooks listed at the beginning for details of them. Again, though, it is most important to
understand the assumptions made and the conceptual picture behind their development.
Reservoir engineers spend most of their time in an exercise called history matching. This is
when a model of the reservoir is reconciled with production data. Although the details can
be very different, conceptually a mathematical representation of the key production process
is tuned to match data and properties of the model determined that give the best match.
Once this is done, predictions of future performance can be made. In this section of the notes,
specifically we will match a model of water influx to production data.
Schematic diagram showing a gas field at initial pressure and saturation (left), and at
sometime later (right), during production, when the pressure has decreased. In this case
we allow for water influx indicated by the blue arrows.
Compared to the case without water influx, aquifer support decreases the reservoir volume
of gas present in the field by We, and the surface volume by We/Bg. Hence, instead of Eq. (4.7)
the remaining gas volume is:
− = −
(4.17)
= 1− +
(4.18)
Note that water influx increases production for a given pressure drop.
We/GBgi is the fraction of the original hydrocarbon pore volume flooded by water. It is always
<1.
As mentioned above, there are several ways in which the aquifer response can be modelled.
The simplest approach is the so-called pot aquifer model: here it is assumed that the gas field
is in contact with an aquifer that responds instantly to any decline in average reservoir
pressure. This is appropriate for relatively slow production, smaller aquifers and well-
connected (high permeability) aquifers.
We assume that the aquifer influx is simply due to the expansion of the rock and the water
that fully saturates the pore space. Let the aquifer have a porosity and gross rock volume
Va; the water volume of the aquifer (measured at reservoir conditions) W=Va. There are two
contributions to the water influx. The first is simply the expansion of water itself: any
expansion of water contributes to the water influx. The second contribution is due to the
compressibility of the rock. Imagine that the pressure drops and amount DP, causing a
Δ . This is the second contribution to the water influx. By analogy with Eq. (4.10),
the change in aquifer volume – which is the water influx – is given by:
= −
(4.19)
The first term is due to water expansion alone at fixed porosity, while the second – and note
the minus sign because the pressure drop represents a decrease in pressure – accounts for
the decrease in porosity with pressure that squeezes water out of the aquifer adding to the
water influx. When the water expands and the pore space compresses, the water has to go
somewhere – it invades the gas field and contributes to the water influx. Then using the
definitions Eqs. (4.11) and (4.13) for compressibility:
=− −
(4.20)
The minus signs indicate that water influx increases as the pressure decreases.
Then, assuming constant compressibilities with pressure, we can write the following for the
water influx. P = (Pi-P) is a positive number, even though the pressure drops – this is why
the minus sign disappears from the equation:11
= + Δ = Δ
(4.21)
where c is the total compressibility of the water and rock. With increasing aquifer size, or rock
compressibility, see larger production for given pressure drop.
With sufficient pressure production history you can determine the gas initially in place if you
have an aquifer model. For our pot aquifer model, the material balance equation, from Eqs.
(4.18) and (4.21):
11
The appearing and then disappearing minus signs can be confusing on first reading. Rather
than puzzle over this, for each equation consider carefully what makes physical sense. The
mathematics has to make the same sense, so this determines what terms are negative and
what are not.
Δ
= 1− +
(4.22)
Δ
= +
1− −
(4.23)
This is an equation of a straight line, where the intercept is G and the slope is Wc (it makes
sense to consider this combined variable the unknown – we do not necessarily need to know
W and c separately).
Given production data (Gp) as a function of pressure P, plus a determination of Bg, you can
determine G and the strength of the aquifer (Wc) as follows. Plot on the y axis as a
function of on the x axis. The intercept (value of y) when x=0 is G, while the slope is
Wc.
Let’s give a simple example from the table below. What is shown in bold is data – that is
information you have to know or determine it before you can perform the analysis. The other
two columns are calculated from the data.
Δ =
= 1−
Gp (million scf) P (MPa) Bg (rb/scf) −
0 34 0.00234
72 33 0.00298 1562.5 335.25
112 32 0.00345 1801.801802 348.1081081
129 31 0.00354 2500 380.55
155 30 0.00399 2424.242424 374.8181818
Table illustrating an aquifer analysis for a gas field as described in the text.
400
y = 0.0466x + 263.05
350
R² = 0.9971
Graph plotting the data from the table above. The slope gives the aquifer strength (Wc)
while the intercept is the original gas in place, G. Be careful with units and do not quote
answers with inappropriate accuracy.
It is straightforward to determine the unknown parameters from the graph – indeed Excel will
give you a best-fit straight line automatically. However, there are two things to pay very
careful attention to. The first is units: G has units (in this example) of MMscf. Wc has the
units of the y axis (MMscf) divided by the units of the x axis (MPa.scf/rb). In their own context
both the MM and M stand for million, so cancel, leaving the units of Wc as rb/Pa. This does
need to be thought through very carefully for every example.
The second common mistake is to assert a ludicrous precision and implied accuracy to your
results. This is often encouraged by the confident quoting of many significant figures for the
best-fit straight line. With real field data, a determination of G to within 10% by this method
is reasonable – in no circumstances would quoting G to more than two significant figures be
appropriate.
Hence, the final result, for this example, should be quoted as: G = 260 MMscf and Wc = 0.047
(or 0.05) rb/Pa.
It is common for an engineer to plot, first, a P/Z plot and then determine if it is a straight line.
A deviation from a straight line is said to indicate the presence of an active aquifer; at this
stage plots with different aquifer models are attempted to find a match. However, this is not
the approach that I recommend. Instead, perform this plot immediately, as deducing an
aquifer from the P/Z plot is often not straightforward and an engineer is often tempted to
assume that it is absent: it is easy to convince yourself that noisy data falls on a straight line.
If there is no aquifer, the graph will be a horizontal straight line – any slope indicates an
aquifer.
For these reasons the graph presented above is a powerful tool in the analysis of a gas field,
enabling the initial gas in place and the strength of the aquifer support to be determined.
The advantage of water influx is that for a given pressure decline there is more production –
this leads to faster recovery, or a reduced need to compress or pump the gas. The
disadvantage is that the water, when it displaces gas, traps residual gas as droplets in the pore
space. This is discussed later in the notes. The residual gas leads to a lower final recovery –
so the gas is produced faster, at a higher pressure, but the cumulative is generally lower.
The residual gas saturation is not known a priori – it is determined from core flood tests on
samples of reservoir rock. Let Sgr be the residual gas saturation.
We will now calculate the water influx necessary to sweep the entire gas field to this residual
– this represents the maximum theoretical recovery; in practice there will be significant water
production before this point is reached. The gas saturation is initially Sgi = 1-Swc and drops to
Sgr. The change in saturation is 1-Swc-Sgr. This is a change in reservoir gas volume of V(1-Swc-
Sgr). Hence this represents a water influx:
= Δ = 1− −
(4.24)
We also know that V(1-Swc) = GBgi, since this is the reservoir-condition initial volume of gas.
Hence:
1− −
Δ =
(1 − )
(4.25)
Eq. (4.24) can be used to find the final pressure at which water influx will sweep the entire
field. If P=Pi-Pf, where Pf is the final pressure and recalling that we have determined G and
Wc from the previous analysis:
1− −
= −
(1 − )
(4.26)
Having found the final pressure, we can then determine the final recovery and recovery factor
from Eq. (4.22):
1− −
= = 1− + =1−
(1 − ) (1 − )
(4.27)
using Eq. (4.25). Note that this equation has the same form as Eq. (4.5) – our first application
of material balance – if we replace the subscript “o” for oil with “g” for gas and note that the
initial saturation is 1-Swc and the final saturation is Sgr. So, this equation can be derived simply
and quickly, direct from a consideration of reservoir and surface volumes: you should be able
to do this readily by now.
There is a subtlety, however: the value of Bg used here is the Bg at the final reservoir pressure,
computed from Eq. (4.26), and not the Bg for the last production data point.
This concludes the discussion of gas fields. Material balance is a particularly effective tool for
analysis in these cases, since there is a significant depletion in pressure accompanied by a non-
linear expansion of gas which can be distinguished from aquifer support. This analysis should
always be performed, either as a complement to a reservoir simulation study, or on its own:
there is never an excuse to ignore this straightforward calculation in favour of some fancy
computer-based analysis.12
12 Even though, of course, there are software packages available to perform material balance
analysis. This is fine, as long as the engineer retains a proper understanding of the results
and approximations used.
Here we will derive and apply the Schilthuis material balance equation (1941) and its
subsequent simplifications. This is the general form of the material balance equation that
accounts for oil, water and gas expansion, aquifer influx, and the compression of rock. We
will use it to study oil fields which may or may not also have a gas cap present.
The conceptual picture is as follows and shown in the diagram over the page. The reservoir
has some pore volume which holds oil, water and gas. The reservoir is described as a tank with
uniform properties that only vary with pressure. As the pressure drops, the fluids expand and
the pore volume decreases as the rock compresses. How is this change in volume
accommodated? The fluids have to go somewhere – they are produced. Thus the change in
reservoir volume of rock and fluids is equal to the reservoir volume of the production. We
construct the material balance equation by going through each contribution to production in
turn: we find the increase in reservoir volume due to expansion. Don’t worry about possible
production – at the end we equate this to the reservoir volume of the fluids produced.
(1 − )
=
(4.28)
where V is the gross rock volume of the portion of the reservoir containing oil.
We define m as the initial hydrocarbon volume in the gas cap divided by the initial volume of
oil. This is measured at reservoir conditions.
Rp is the cumulative gas/oil ratio. It is the cumulative gas production divided by the cumulative
oil production. This is measured at standard conditions. Note that Rp is distinct from the
solution gas/oil ratio Rs and the producing gas/oil ratio R (which measures the ratio of the
rates of production, not the cumulative).
Expansion of solution
and gas cap gas
Compression of the
pore space
Water influx
A diagrammatic representation of material balance. The reservoir is treated as a tank
with uniform properties. The reservoir volume of the expansion of rock and fluids is
equal to the reservoir volume of the fluids produced.
Initially, through these definitions, at reservoir conditions, the oil volume is NBoi and the gas
cap volume is mNBoi. The surface volume of the gas cap initially, defined as G in the previous
section, is therefore:
= /
(4.29)
We now consider the expansion – measured at reservoir conditions – of the fluids and rock
after a period of production.
Expansion of oil and solution gas Eo. The expansion of liquid oil alone is the volume of oil now
minus the initial volume, N(Bo-Boi). This term can be negative – the oil volume may shrink
because of the exsolution of gas. However, in this case there is another contribution, which is
the expansion of gas liberated from solution. At initial conditions, if all the oil were brought
to the surface, the volume of solution gas would be – by definition – NRsi. At some later time
this volume would be NRs, and would be a smaller volume if we are below the bubble point,
because some gas has already come out of solution. The difference, N(Rsi-Rs), is the surface
volume of solution gas that has exsolved. We convert this to a reservoir volume using the gas
formation volume factor; hence the contribution of solution gas to expansion is NBg(Rsi-Rs).
The total expansion of oil and associated gas is therefore:
= − + ( − )
(4.30)
This term is always positive – overall the combination of oil and solution gas expands as the
pressure drops. Note that Eo alone is the fractional change in volume; for the change itself
you need to multiply by N.
Gas cap expansion, Eg. The initial surface volume of the gas cap is G, given by Eq. (4.29). The
initial reservoir volume of the gas cap is hence GBgi. Sometime later, the surface volume of
this gas is the same, G (remember we do not account for production here): this has a reservoir
volume GBg. The change in reservoir volume of gas – similar to that of oil – is then G(Bg-Bgi).
Applying Eq. (4.29) we find:
= −1
(4.31)
Note that like Eo, Eg is the fractional change in volume – it is multiplied by mN to find the actual
change in gas volume.
Initial water expansion and rock compressibility, Er. These effects have already been
considered in the context of gas reservoirs. Let Vh be the total hydrocarbon pore volume (oil
and gas) and Vt to be the gross rock volume of the reservoir – including both the oil column
and the gas cap. We write for the hydrocarbon volume Vh=Vt(1-Swc) = Vt - Vw, where Vt is
the gross rock volume of the oil field plus gas cap and Vw is the total water volume in the oil
field plus gas cap. Then, similarly to Eq. (4.10) we can write:
= −
(4.32)
As before, the first term on the right-hand-side of Eq. (4.32) represents the rock
compressibility and the second the expansion of water. We apply the water and rock
compressibilities given by Eqs. (4.12) and (4.13) respectively to find – similar to Eq. (4.14):
1 +
=
1−
(4.33)
Then for constant compressibilities and defining, as before, P=Pi-P (and being very careful
over signs – again use physical intuition to tell you what is correct and then the mathematics
will look after itself):
+
Δ = Δ
1−
+
= (1 + ) Δ
1−
(4.34)
using the earlier definitions to state Vh=(1+m)NBoi. This change in reservoir pore volume of
hydrocarbon also contributes to production and we define this expansion from Eq. (4.24) as:
+
(1 + ) = (1 + ) Δ
1−
(4.35)
Water influx, We. This has also already been discussed in the context of gas reservoirs. In
general, this is simply left as the term We and later different putative aquifer models are fitted
to the data. In these notes we consider only the pot aquifer model with We = WcP, Eq. (4.21).
Total production, F. At surface conditions we have produced Np oil, Gp=RpNp gas (defining Rp)
and Wp water (again this defines Wp). We need to convert these volumes to reservoir
conditions using the relevant formation volume factors. For the gas, we need to remove the
amount that dissolves in oil – this is RsNp. Hence:
= +( − ) +
(4.36)
Notice the introduction of the water formation volume factor Bw. This is generally close to 1,
but may be larger thanks to the presence of dissolved gases in the brine (such as carbon
dioxide) at reservoir conditions.
Final material balance equation. The material balance equation is now constructed by
making the reservoir volume of produced fluids equal to the reservoir volume of the different
expansion terms:
= ( + + (1 + ) )+
Or, in full:
+( − ) +
= − + ( − ) + −1
+
+ (1 + ) Δ + Δ
1−
(4.37)
The advantage of the material balance analysis is that it is a simple, quick application to real
reservoirs, knowing only the phase behaviour and production history as functions of pressure.
It provides valuable insight into the size of the connected reservoir and the production
process.
The major disadvantages of the method are the lack of time dependence, and that the
properties are reservoir averaged. Reservoir simulation on large fields often does a much
better job, particularly if fluids are injected to maintain the pressure. Material balance may
also make poor or uncertain predictions is there is not much history.
We will now use the material balance equation to study recovery for different reservoir
production mechanisms: material balance can be used to determine the relative contribution
to recovery of different production processes. In theory, we can take the full material balance
equation (4.37) and find the best match to the three unknowns: N, m and Wc. We can also
determine the relative sizes of each of the expansion terms contributing to production. This
is routinely performed by software; here, instead, we take a simpler approach and will
consider examples when one mechanism dominates and we have only two unknown
parameters to find, which is achieved by plotting the data as a straight line.
Here we assume that there is no gas cap initially present. In this case the production is caused
by the expansion of oil and solution gas plus water influx.
Firstly, how might we suspect that there is no gas cap? Consider the initial reservoir pressure
– particularly near the top of the formation. If a gas cap is present, the oil pressure in contact
with the gas must be the bubble point pressure; the gas pressure must be the dew point. Why
is this? The oil and gas have remained in thermodynamic equilibrium for millions of years. If
the oil were at a pressure above the bubble point, gas would dissolve in the oil until the bubble
point was reached. Any excess gas would migrate upwards to the gas cap. Similarly, the gas
is at the dew point, since if it were at a higher pressure, oil would dissolve and any excess oil
migrates downwards to the oil column. The only way for gas and oil to be in mutual
equilibrium is as described. Hence, an oil whose initial pressure is well above the bubble point
is unlikely to have a connected gas cap – any gas in the field is unlikely to be in contact with
the oil. What is meant by “well above”. Well, first consider the uncertainty in the
measurement or prediction of the bubble point. Also, the datum depth may not be the top of
the formation, and you need to correct for this (we discussed how pressure varies with depth
previously in these notes). If, even accounting for this, it is highly unlikely that the oil pressure
at the gas/oil contact can be at the bubble point, then it is also highly unlikely that there is a
gas cap. The converse though is not necessarily the case – if the oil pressure is close to the
bubble point pressure there is not necessarily a gas cap present; this could simply be
coincidence.
The second inference for a gas cap comes from direct log measurements – is a gas/oil contact
detected – and seismic, which may be able to infer the presence of gas directly. A good
engineer combines information and develops a model consistent with all the data. Trying to
match production data with a gas cap because this is more “general” or “accurate” is senseless
if there is other clear evidence that a gas cap is absent.
Determining the likely presence of an aquifer is more difficult, as most fields have an oil
column in direct contact with water-saturated rock. Of course, if this is not the case, there is
not an aquifer and trying to assume one is again a pointless exercise. However, the presence
of water in contact with oil does not prove an aquifer in an engineering sense: how large is
the connected aquifer and is there sufficient permeability to make a significant difference to
production? This is often difficult, if not impossible, to determine definitively from log and
seismic information alone. Having a sophisticated simulation model does not help here either,
since you have to input the aquifer in the first place – the behaviour is determined by the
assumptions you make and does not magically emerge from the model. The sensible use of
reservoir analogues may be some guide; well test analysis may be able to determine the
locations of reservoir boundaries to infer or preclude the presence of an aquifer. In the end
though, it is often a material balance analysis that is used to determine if an aquifer is present
and its strength – something that we will demonstrate below.
= ( + )+
+
+ = ( − )+ Δ + Δ
1−
(4.38)
The first term – the oil expansion – can be written in terms of oil compressibility using Eq.
(4.11). N(Bo-Boi) is the change in the reservoir volume of oil. Assuming a constant
compressibility we can write:
−
= Δ
(4.39)
+
+ = + + Δ
1−
(4.40)
While this looks complicated, physically Eq. (4.40) states that the reservoir volume of
produced fluids is proportional to the pressure drop. Hence, the production simply scales with
the pressure decline. But what is the constant of proportionality? This is the weighted
contribution of the compressibility of oil, water, rock and the aquifer, which is constant with
pressure, to a good approximation. While rock, water and oil compressibility can be
measured, the relative contribution of the aquifer compared to the oil cannot be determined
from this analysis alone, since we do not know W and N independently. Hence, in this case,
material balance fails to distinguish between strong production from a large oil field (and
relatively little aquifer support) or a small field and a large aquifer. Additional data – such as
Again, it is more important to understand this concept than rely on software output that might
find a “best match” but one with very little real accuracy or confidence.
If there is no aquifer support and we ignore water production, then Eq. (4.40) can be further
simplified to:
= Δ
(4.41)
+
= +
(4.42)
Using example values cw = 0.5×10-9 Pa-1, cf = 10-9 Pa-1, co = 1.5×10-9 Pa-1, Swc = 0.2 and Soi = 0.8,
we find an effective compressibility ce ≈ 3×10-9 Pa-1.
Then, for instance, we can calculate the recovery at the bubble point:
= = Δ
(4.43)
If we take a case where the initial reservoir pressure is 100 atm (10 7 Pa) above the bubble
point, then the recovery from Eq. (4.43) is around 3%, assuming a relatively small change in
Bo with pressure.
This analysis indicates that recovery factors from reservoirs without a gas cap or significant
aquifer are only a few percent, if the pressure is maintained above the bubble point: to
achieve significant recovery, water (or gas) must be injected to maintain pressure and displace
the oil.
development. One key uncertainty in reservoir development is often whether or not there is
an active aquifer that could contribute significantly to production. Unfortunately, investment
decisions are frequently made based on optimistic hunches or experience with “similar” fields
which often turn out to be incorrect. It is always valuable to perform a worst case scenario –
one with no aquifer – and allow flexibility in the design of production facilities to allow for
water injection in this case. How long will it be before water injection is required to maintain
pressure? The evasive answer is you can’t tell, because material balance does not include
time dependence. But now – like any good reservoir engineer – additional information needs
to be incorporated. Normally facilities are designed with a target production rate – say Q (in
stb/day). There will also be an estimate (or better still a range) of initial oil in place N. So the
time to reach the bubble point is simply NRf/Q (in days), where Rf is given by Eq. (4.43).
Normally you expect to produce around 2% of the field (at least) per year, so the time is
generally around 1-2 years. You need to ensure that if the pressure does indeed decline as
predicted from the no-aquifer case, it is possible to incorporate water injection facilities in
time, otherwise you may be constrained on production rate, or fall below the bubble point
and damage the reservoir by introducing a gas phase. The same approach can also be used –
as shown later – if there is an aquifer model. With the target production rate, how rapidly do
you predict the reservoir pressure to decline and do you have a plan in place to maintain
pressure if needed?
Now consider that we drop the pressure below the bubble point. Here recovery is dominated
by the expansion of gas liberated from solution – this is called a solution gas drive. As we have
shown in the context of gas reservoirs, the gas compressibility is typically many times greater
than that of water and rock and so – for simplicity – it is reasonable to ignore the rock and
connate water terms in the material balance analysis. They can always be included if there is
good data for rock compressibility, or if this compressibility is high (say a poorly consolidated
formation).
If we assume no water influx (let us take a scenario where the observed rapid decline in
pressure to the bubble point has precluded the existence of an active aquifer) then Eq. (4.37)
reduces to:
+( − ) = − + ( − )
(4.44)
The way to identify a solution gas drive is that the ratio F/Eo is a constant with pressure – its
value is the initial oil in place, N: later we show a better way to analyse this including the
possible presence of a gas cap. The recovery factor is given by:
− + ( − )
= =
+( − )
(4.45)
The expansion of gas contributes significantly to recovery until the gas saturation reaches
some critical value, Sgc, at which point the gas is connected through the pore space and is
produced preferentially to oil. Then excessive quantities of gas are produced and the field
essentially becomes a gas field, leaving the vast majority of the more valuable oil behind. In
principle Sgc can be measured in the laboratory, although defining a representative value is
challenging: often, mistakenly, in simulation models a high value of Sgc is assumed below
which the gas relative permeability is set to zero. This, by construction allows the pressure to
drop below the bubble point for a time without any gas production: the reality is that there is
some gas flow even at very low saturation, but this is significant only when there is good
connectivity of the gas. An accurate model of the gas relative permeability is required and
this is rarely achieved through the simplistic assignment of an optimistic critical gas saturation
(it is often better to set it to zero in simulation models); relative permeability is discussed
further later in these notes.
The average gas saturation in the field can be estimated from material balance once the value
of N has been determined. The initial volume of oil in the reservoir is NBoi (measured at
reservoir conditions). The oil volume sometime later, during production is (N-Np)Bo (the oil
that has not been produced converted to reservoir conditions at the current pressure). If we
ignore changes in the pore volume from rock compression and connate water expansion (this
is a relatively small effect in the presence of gas), then the change in volume is accommodated
by gas. We convert this to a gas saturation by dividing by the pore volume NBoi/(1-Swc). Hence
we find:13
−( − )
= = 1− 1− (1 − )
/(1 − )
(4.46)
In terms of saturation, we can rearrange Eq. (4.46) for the recovery factor:
1− −
= =1−
1−
(4.47)
This is a complex way to derive the expression that can be obtained directly from material
balance – see Eq. (4.5)
Let’s take a typical example with Sg = Sgc = 0.3, Swc = 0.2 and Bo/Boi = 0.88. This gives Rf = 29%.
Recoveries from a solution gas drive are generally in the 20-30% range before the field has to
be abandoned due to excessive gas production. Furthermore, it is then difficult to recover
more oil from the field: water injection below the bubble point requires an increase in
pressure to boost recovery, so the water simply compresses gas and is initially very inefficient
in recovering additional oil.
In modern reservoir engineering, a solution gas drive is only a good option as a tertiary
recovery mechanism. A field is produced initially under primary production to the bubble
point. Then it is waterflooded and pressure is maintained. Then, and only then, when the
field is mainly residual oil, is the pressure dropped. This is achieved simply and cheaply by
producing oil and not injecting water. The field is then essentially managed as a gas field to
liberate the solution gas. This is a good option for light oils where there is a network to sell
the produced gas. This concept was pioneered by Shell for the Brent field in the North Sea
and is now frequently considered for mature fields in provinces with a good gas pipeline
infrastructure.
13 Note that Dake on p86 gives the wrong expression for gas saturation.
Here we consider expansion of a gas cap, but we will ignore water influx and the (relatively
small) contribution of rock compressibility and connate water expansion. As mentioned
previously, the initial reservoir pressure is the bubble point and, as production proceeds, the
pressure drops below this. As a result there is recovery due both to the expansion of gas in
the gas cap itself, and solution gas. The material balance equation, (4.37) becomes:
= ( + )
+( − ) = − + ( − )+ −1
(4.48)
/ = + /
(4.49)
So we take the production data and fluid properties as a function of pressure. We compute
F, Eo and Eg for each pressure value. We then plot F/Eo on the y axis and Eg/Eo on the x axis.14
The intercept when x=0 is N, the original oil in place. A slope to the data plotted this way
indicates the presence of a gas cap (if there is no slope we have a pure solution gas drive and
there is no active gas cap). The slope itself has a value Nm, from which the relative gas cap
size, m, can be easily found. Here the units should be straightforward – even if field units are
used, no conversions are needed, as we are dealing in ratios of quantities. N is measured in
stb while m is a dimensionless ratio.
For clarity, let’s show a simple example. As for the gas reservoir case the data – information
you need to perform the analysis – is shown in bold, while the other quantities are calculated
from the data.
14 There is no point for the initial condition, as the ratios are 0/0.
Table illustrating an example material balance analysis for an oilfield with no aquifer but
the possible presence of a gas cap, as described in the text.
The graph is plotted as shown overleaf. The best fit to the data is N=250 MMstb and the slope
of the line gives mN = 7.6 MMstb, or m = 0.03. In this case we have a relatively small gas cap
(only 3% of the total oil volume), yet the expansion of gas is considerably larger than the
expansion of oil. There is a large expansion of the relatively small volume of gas in the gas
cap: recall though that the overall recovery at this stage of production is still very small (only
around 2%).
1000
Volume of fluid produced/epxansion of oil,
500
400
300
200
100
0
0 20 40 60 80 100
Expansion of gas/expansion of oil, Eg/Eo
Graph plotting the data from the table above. The intercept gives N, the value of the
initial oil in place, while the slope is mN where m is the relative size of the gas cap.
However, the best way to understand this is to perform the exercise yourself: at the end of
the notes there are many exam questions that can be used for practice. Once the mechanics
of the analysis are known, you can use software and apply the methods to real field cases with
some confidence. My recommendation is to use this plot, even if you are not sure about the
presence of a gas cap – if there is none, then the points simply lie on a horizontal line.
The presence of a gas cap can allow higher recoveries than solution gas alone: the gas is at the
top of the reservoir and when it expands, it pushes the oil downwards. With wise well
completions near the base of the oil column (or horizontal wells low in the oil column) the
problem of excessive gas production can be mitigated, although withdrawing at a high rate
encourages coning and an increase in the gas/oil ratio. Typical recoveries are 25 - 35%, but
values as high as 70% are possible for very slow production, allowing the gas to displace oil
down to very low final saturations. The mechanism for this – oil layer drainage – is discussed
later in the notes. In our example however, with a very small gas cap, the recovery is very
modest (only around 2%) for a significant drop in pressure – here we have principally the
production of solution gas from the reservoir.
Natural water drive, or aquifer drive, has already been discussed in the context of gas fields.
Here, for simplicity, we assume that there is no gas cap and we assume that the effects of rock
compressibility and connate water expansion are relatively small. The material balance
equation (4.37) becomes:
= + Δ
+( − ) + = − + ( − ) + Δ
(4.50)
Here we have assumed a simple pot aquifer model, although others can be used to match the
data more accurately, where needed.
/ = + Δ /
(4.51)
As for a gas cap drive, we plot F/Eo on the y axis. On the x axis we plot P/Eo. The intercept
where y=0 is, again, N the initial oil in place. The slope is Wc. Here, as with the analysis of a
gas reservoir, care needs to be taken with the assignment of units. In field units Eo is given in
rb/stb. If P is reported in psi, then Wc has units of rb/psi. Once again, the best way to
understand this is through working through some of the example exam questions – in this
case I will not give a worked example.
The final primary recovery mechanism is compaction drive, where the rock compressibility is
significant. An example is the Bachaquero field in Venezuela where c = 15×10-9 Pa-1 and
compaction accounts for 50% of oil recovery; rock compression is also significant, for instance,
for the poorly consolidated and unconsolidated sandstones in the Gulf of Mexico.
If rock compression dominates over aquifer support (and connate water expansion) with no
gas cap, we can write for Eq. (4.37):
= ( + )
+( − ) + = − + ( − ) + Δ /
(4.52)
By this stage, you should see the approach to use: once more we plot F/Eo on the y axis. On
the x axis we plot BoiP/Eo. The intercept is N, as before, while the slope is Nc/Soi from which
the rock compressibility can be found:
= + Δ /
(4.53)
Of course, it is possible to measure rock compressibility from rock samples taken from the
field; however, this is not necessarily representative of the field as a whole. The advantage of
this method is that production data is used to deduce a flow-averaged compressibility – this
should be compared with the measured values for consistency, but it is the value from
material balance that is the more robust, as it represents the average behaviour of the whole
field under production.
This now concludes the analysis of different primary recovery production mechanisms using
the material balance analysis. As said previously, the data can usually be analysed using
commercial software; this is fine and allows the consideration of more than one production
mechanism simultaneously, but is no excuse for not having a clear understanding of the
production process yourself.
The material balance analysis does not take rate – fluid flow – directly into account. It
therefore assumes that the recovery behaviour is insensitive to the rate at which the field is
being produced. This is a reasonable approximation for solution gas drives (dominated by gas
expansion and evolution) and a strong water drive with a well-connected aquifer.
However, reservoirs above the bubble point with a weak aquifer drive are sensitive to the
production rate. If the pressure is dropped very rapidly, then the aquifer cannot expand
sufficiently quickly in response, and so the recovery is lower (for a given pressure drop) than
a slower production that allows the aquifer time to move into the oil reservoir.
Other production processes that are rate sensitive include any situation where gravity
segregation is important (for instance where gas evolves and rises to the top of the formation).
Rapid production does not allow sufficient time for the fluids to segregate, resulting in poorer
production (in general) than cases where the fluids do separate. This is also the case where
coning is significant – higher production rates lead to more coning and increased production
of water and/or gas.
This needs to be borne in mind when designing target production rates – in the end though a
reservoir simulation approach is needed to assess the sensitivity of recovery to rate.
While a material balance analysis should be performed for any field with a substantial
pressure drop, it is the main method of analysis for primary production data for small fields,
including most gas fields.
Production history is used to predict oil and gas reserves and ultimate recovery. The more
production there has been, the more accurate the predictions.
The approach is conceptually similar to that seen in other areas of petroleum engineering,
such as well test analysis. The first – and often most difficult step – is to decide on the
reservoir drive process (model identification). For this the engineer should incorporate all the
data that is available to make a sensible assessment of the likely production mechanism.
In these notes, I have shown how to plot the data so that it lies on a straight line, with the
slope and intercept giving you the required parameters. Once again, to understand this more
fully, I recommend attempting some of the homework problems given at the end of these
notes.
Decline curve analysis is an empirical way to forecast production decline from measurements
of production rate; it acts as a simple complement to material balance analysis, where rate
dependence is not included.
The method is important for project economics. The rate of decline will depend on the
wellbore condition, well spacing, surface facilities, porosity, permeability, reservoir thickness,
fractures, relative permeability, formation damage, drive mechanism, compressibility and gas
production. Ideally – and indeed as shown later in these notes – the production data should
be matched to a physically-based flow model (either analytical or simulation). Here briefly,
we simply introduce the nomenclature and some example types of decline. However, this is
not a replacement for a more rigorous approach based on flow modelling.
Most of the equations presented here were first proposed by Arps (1956) before the advent
of reservoir simulation and a modern approach to understanding flow processes. Hence,
while it provides a simple analysis of production, its predictions are highly unreliable unless
associated with a proper model of flow. Decline curve analysis is currently used to interpret
and predict the behaviour of shale oil and gas reservoirs – this is often a reflection that the
flow processes are not fully understood and such analyses are of dubious accuracy at best –
and need to be rooted in an understanding of flow processes.
The analysis assumes that you have stable operation at capacity. Below capacity, no decline
may be seen at all (plateau).
1
=−
(5.1)
where Q is the oil production rate (stb/day) from either the entire field, or a single well – the
analysis can be performed on either.
This is the simplest type of decline where b is assumed to be constant. In this case it should
be obvious that the production rate is given by:
( )=
(5.2)
How do we check if we have exponential decline? Either plot the computed value of b from
Eq. (5.1) against time – or better (since it removes the need to compute numerical derivatives
from the data) – check if lnQ varies linearly with time, t. If so, the slope is –b.
−
= = =
(5.3)
The cumulative production varies linearly with flow rate: this is another check for exponential
decline.
1 /
=− =c
(5.4)
for some constant a. Here the analysis is rather more cumbersome. It is common simply to
compute a decline rate and assume it constant (exponential decline as the default in the
=−
(5.5)
and hence:
/
= +
(5.6)
/
/ / /
= − = −1
(5.7)
/
/
= +1
(5.8)
= =
/
1+ 1+
(5.9)
= =
1+
(5.10)
= 1− 1+ = − 1+
−1 ( − 1)
(5.11)
Strictly hyperbolic decline is a=2 and occurs for gravity drainage and in gas reservoirs. a=1/2
is seen in the early production behaviour of shale gas and fractured reservoirs. Indeed, in
unconventional reservoirs decline curve analysis, in the absence of a more sophisticated
understanding or model of the reservoir, it is often used to assess future production. You
should be able to demonstrate this later using the methods developed in these notes.
Harmonic decline is a special case when a=1. This is observed for high viscosity oil and a water
drive. It may also be seen for high WOR (water/oil ratio) and constant fluid rate, such as in
thermal projects and steam soak.
=
1+
(5.12)
and:
= ln(1 + )
(5.13)
The treatment here of decline curve analysis has been deliberately very brief; it is an empirical
approach which should, where possible, be substantiated with numerical simulation or
analytical results. However, for this we need to understand fluid flow which is the subject of
the subsequent sections of these notes.
We now divert our attention away from oil fields and petroleum recovery, to the details of
the science of how fluids are arranged in the pore space of rocks – at the micron scale. This is
necessary to have a good understanding of how multiple fluids – oil, water and gas – are
configured in the pore space of the rock and how they flow.
We start with a presentation of the fundamental equations which govern contacts between
fluids and solids and the meniscus that separates fluid phases.
If we have multiple phases present in a porous medium, then there is a pressure difference
across the interfaces between these phases. The pressure difference is given by the Young-
Laplace equation:
1 1
= +
(6.1)
where r1 and r2 are the principal radii of curvature of the fluid interface (the radii of curvature
measured perpendicular to each other. is the interfacial tension between the phases – it
has the units of a force per unit length (N/m) or an energy per unit area (J/m2).15
It is possible to derive this equation using principles of force or energy balance (see Dullien,
de Gennes or Blunt – the books listed at the beginning of the notes), but is somewhat
cumbersome, as it involves some obscure mathematical results concerning the three-
dimensional geometry of curved surfaces. It is, however, relatively straightforward to derive
specific cases – as below – from first principles. This we will do later when we consider the
15
Sometimes can be called the surface tension. Strictly this is only valid for a liquid in
equilibrium with its vapour. This is a situation that we will rarely encounter in real situations,
so it is always preferable to refer to the interfacial tension – the energy per unit area of a
boundary between two phases, regardless of composition.
pressure difference between two phases in a circular cylindrical tube. For now, though, it is
helpful to assert that the Young-Laplace equation is valid.
The second major concept we need to introduce is that of contact angle. While the Young-
Laplace equation considers the pressure across an interface, it does not address how that
interface interacts with a solid surface.
Consider a wetting fluid (say water) resting on a solid surface surrounded by a non-wetting
phase (such as oil or gas).
ow
Non-wetting
Wetting
so sw
Solid
Contact angle, θ, between two phases, measured through the denser phase.
(a) non-wetting liquid. (b) wetting liquid. (c) intermediate-wet liquid.
= +
(6.2)
−
=
(6.3)
What is the vertical force balance? Intermolecular forces in the solid counter-act the vertical
tension as the solid is very slightly perturbed.
Thomas Young was an early 19th century English physicist and all-round genius who worked
on everything from deciphering hieroglyphics to the wave theory of light. Young’s modulus
(in elasticity) and the Young double slit experiment recognise just two of his contributions to
science.
Pierre-Simon (the so-called Marquis de) Laplace was a brilliant French mathematician and
physicist of the same era. His contributions include astronomy, statistics and the Laplace
called the ‘Laplace equation’ which is confusing; and Eq. (6.2) the ‘Young-Laplace equation,’
which possibly under-states the relative contribution of Young to this subject.
Spot the difference. One is an insightful English physicist; the other a brilliant French
mathematician, clinging to a post-revolutionary title.
= − −
(6.4)
If so>sw +ow or Cs > 0, there is no solution for in Eq. (6.4). The water spreads over the
solid surface. This is complete wetting and = 0.
2r
Non-wetting R Wetting
Two fluids in equilibrium in a cylindrical tube. The blue arc is the interface between the
wetting and non-wetting phases.
The radius of curvature R = r/cos(we can derive this using simple trigonometry).
Thus we find:
2
= − =
(6.5)
Air
h
Water
2
ℎ=
(6.6)
What would happen for mercury/air? The interface would be lower than the free (flat)
surface, as in this case mercury is the non-wetting phase.
Eq. (6.6) can be derived without the Young-Laplace equation, using an energy balance. Water
rises up the tube, since this is energetically favourable: there is a lower energy for water to
coat the solid (glass) surface than air. This though is balanced by potential energy, as the
water rises against gravity. So, the water continues to rise until the change in potential energy
is just balanced by the change in interfacial energy.
Imagine that the height of the interface changes by a small amount: if this lowers the energy,
then the interface will move until it reaches a position of equilibrium. The equilibrium height
is found when changes in potential and interfacial energy are equal for small fluctuations in
this height: this is the stable configuration.
The gravitational potential energy of a mass m at a height h is mgh. From a height h to h+dh,
the water in the tube has a mass Adh, where A is the area (r2). Then if we change the height
from h to h+dh, the change in potential energy is:
∆ = ℎ ℎ
(6.7)
This must be equal to the energy gained when the water moves up the tube. Now consider
that sa is the energy per unit area (interfacial tension) of an interface between the solid and
air, while sw is the energy per unit area of interface between solid and water. Then the energy
change for the water to move from h to h+dh is:
∆ =2 ( − ) ℎ=2 ℎ
(6.8)
since 2rh is the area of wetted solid (glass) surface, and we have used the Young equation,
(6.2). Conservation of energy asserts that the energies in Eqs. (6.7) and (6.8) are the same,
leading to Eq. (6.6) directly:
∆ =∆
2 ℎ= ℎ ℎ
2 = ℎ
2
= ℎ
(6.9)
6.5 WETTABILITY
The contact angle is traditionally measured through the denser phase (water for oil/water and
gas/water, and oil for gas/oil). If:
Clean rocks are generally water-wet, since the polar surface of the solid – say quartz or calcite
– interacts strongly with the water making its interfacial tension lower than the tension
(energy per unit area) with oil or gas, which has less interaction with the surface.
Why then are most reservoir rocks not completely water-wet? In contact with a solid surface,
surface active components of the oil – high molecular weight molecules called asphaltenes –
adhere to the solid surface rendering it less water-wet. Regions of the solid surface that are
not directly contacted by oil remain water-wet. Thus many oil reservoirs are what is known
as mixed-wet or fractionally wet – different regions of the pore space have different
wettabilities. This important concept is developed later in these notes: the wettability (or
contact angle) is not a constant throughout the rock and depends precisely on the mineralogy
of the surface, surface roughness, the oil and brine compositions and the temperature and
pressure of the reservoir.
In other settings – aquifers and soils – it is the presence of organic material, particularly
surfactants or other compounds that adhere to the solid surface – that alter the wettability.
In general, the contact angle is governed by a subtle balance of surface forces between the
fluids and the rock.
The wettability change typically takes around 1,000 hours to complete, so since oil has been
in a reservoir over geological times, there is plenty of time for the wettability alteration to
occur. The wettability alteration also has sufficient time to take place in most polluted soils.
Overleaf is a schematic of what happens in a single pore with an idealized triangular cross-
section: regions of the surface directly contacted by oil have an altered wettability (this is not
necessarily oil-wet, but is not strongly water-wet either) while the corners remain water-wet.
The contact angle is also usually different depending on the flow direction. The static contact
angle (no movement of the solid/wetting/non-wetting contact), the advancing contact angle
(denser phase advancing) and the receding contact angle may all be different.
The three reasons for this are wettability alteration (described above), chemical
inhomogeneities on the surface, and small-scale surface roughness.
Oil
Wate
Oil and water in a triangular pore after primary drainage (oil migration into the reservoir).
The areas directly contacted by oil (shown by the bold line) have an altered wettability,
while the corners that are water-filled remain water-wet. b is the length of the water-wet
surface.
The next figure shows how the advancing and receding contact values vary with the intrinsic
contact angle (the angle measured at rest on a smooth surface).
Displacement – the movement of one phase across another – is always impeded by the most
difficult step, or part, of the motion. This is an important concept that we will meet many
times in these notes. So, if we imagine a surface that is chemically heterogeneous with water-
wet and oil-wet patches, then water will invade the water-wet regions quickly and get
impeded by oil-wet portions. The pressure – the highest pressure – in the water necessary to
move across these patches will be the same as if the system were entirely oil-wet – hence the
system looks to water as if it is oil-wet. Now consider the reverse: oil invasion. In this case,
movement across the oil patches is easy (occurs at a low oil pressure) but is impeded by the
water-wet regions. Hence, for oil invasion, the surface appears water-wet. In terms of contact
angle, the (water) advancing angle is greater than 90o, while the receding angle is less than
90o: we see significant contact angle hysteresis.
Advancing
Receding
CO2
Brine
Rock surface
Relationship between intrinsic contact angle (the angle at rest on a smooth surface) and
the advancing and receding contact angles, based on the measurements by Morrow
(1975). The principal reason for this contact angle hysteresis is surface roughness – in
a porous medium, the solid surfaces are not smooth.
The main effect though in porous media is usually small-scale surface roughness. The previous
figure shows the measured relationship between advancing and receding contact angle on a
rough surface while the effect is shown schematically on the next page. While the molecular-
scale contact angle may indeed be the intrinsic value, the apparent angle viewed on the larger
scale – and the angle which will control the capillary pressure (that is the interfacial curvature)
for displacement is different. As mentioned previously, displacement is impeded by the
highest pressure in the displacing (advancing) phase necessary for the contact to move: hence
the apparent angles are different in imbibition (wetting phase advancing) than drainage (non-
wetting phase advancing). The interface is impeded at the points where the pressure in the
advancing phase is largest – we need to capture this to calculate the correct capillary pressures
for displacement.
Non-wettting phase a
Wettting phase
Wettting phase advance (imbibition)
i
A schematic of surface roughness (the solid surface is indicated by the irregular brown
line) and its impact on contact angle. i is the intrinsic contact angle – the angle at the
surface (regardless of orientation): in this example it is close to zero, indicating a
strongly water-wet system. However, the effective dynamic contact angle for imbibition
a is larger – this is the angle that gives the correct curvature in the Young-Laplace
equation and is observed at a larger scale with an apparently smoother solid surface
(shown by the dashed line). The roughness impedes imbibition: a higher wetting phase
pressure is required for displacement than on a smooth surface. For drainage the
displacement is limited by roughness oriented in the other direction and the apparent
contact angle is smaller than I, impeding the advance of the non-wetting phase.
7. POROUS MEDIA
We now properly introduce multiple phases in a porous medium. Before proceeding with the
description of macroscopic, averaged, properties, we will first use some illustrative examples
in an attempt to provide some insight into displacement and fluid configurations at the pore
scale. The emphasis in this section will be on understanding and appreciating the complexity
of the pore space at the micron (pore) scale.
In recent years there has been a revolution in our ability to see inside the pore space of rocks
using X-rays. The development of modern imaging methods relies on the acquisition of three-
dimensional reconstructions from a series of two-dimensional projections taken at different
angles: the sample is rotated and the adsorption of the X-rays in different directions is
recorded and used to produce a three-dimensional representation of the rock and fluids. In
the 1980s these methods were first applied in laboratory-based systems to measure two and
three-phase fluid saturations for soil science and petroleum applications with a resolution of
around 1-3 mm. The first micro-CT (micron or pore-scale) images of rocks were obtained by
Flannery and co-workers at Exxon Research using both laboratory and synchrotron sources
(Flannery et al., 1987). In a synchrotron a bright monochromatic beam of X-rays is shone
through a small rock sample. Several rocks were studied with resolutions down to around 3
m. Dunsmuir et al. (1991) extended this work to characterize pore space topology and
transport in sandstones.
One of the pioneers of the continued development of this technology has been the team at
the Australian National University in collaboration with colleagues at the University of New
South Wales (see, for instance, Arns et al., 2001, 2005 ). They have built a bespoke laboratory
facility to image a wide variety of rock samples and then predict flow properties: this work is
also now available as a commercial service. The base image is a three-dimensional map of X-
ray adsorption; this is thresholded to elucidate different mineralogies, clays and, principally,
to distinguish grain from pore space.
of our instrument at Imperial College is shown on the front cover of these notes together with
the core holder into which a small cylinder of rock is fitted.
In a micro-CT scanner the X-rays are polychromatic and the beam is not collimated – the image
resolution is determined primarily by the proximity of the rock sample to the source. These
machines offer the advantage that access to central synchrotron facilities or a custom-
designed laboratory is not required, and there is no constraint on the time taken to acquire
the image, allowing signal to noise to be improved. The disadvantage is that the intensity of
the X-rays is poor compared to synchrotrons while the spreading of the beam and the range
of wavelengths introduces imaging artefacts.
The figure on the next page shows two-dimensional cross-sections of three-dimensional grey-
scale images for eight representative rock samples: several carbonates, including a reservoir
sample, a sandstone and a sand pack. The images were acquired either with a synchrotron
beamline (SYRMEP beamline at the ELETTRA synchrotron in Trieste, Italy) or from a micro-CT
instrument (Xradia Versa).
For the quarry carbonates shown in the figure (Estaillades, Ketton, Portland, Guiting and
Mount Gambier) a connected pore space is resolved, although the details of the structure are
complex and at least two of the samples – Ketton and Guiting – are likely to contain significant
micro-porosity that is not captured with the resolution of the image. Also included is a
carbonate from a Middle Eastern aquifer. In this case, while some pores are shown with a
voxel size of almost 8 m, it is likely that there is significant connectivity provided by pores
that are below the resolution of the image.
I now show, overleaf, example three-dimensional images of three carbonates where only the
pore space is shown. Ketton is a classic oolitic limestone composed of almost spherical grains
with large, well-connected pores between them. Estaillades has a much more complex
structure with some very fine features that may not be fully captured by the image. Mount
Gambier has a very irregular pore space, but it is well connected and the porosity and
permeability are very high. Overall, while a resolution of a few microns can resolve the pore
space for some permeable sandstones and carbonates, many carbonates and unconventional
sources, such as shales, contain voids that have typical sizes of much less than a micron.
Typical X-ray energies are in the range 30-160 keV for micro-CT machines – with
corresponding wavelengths 0.04-0.01 nm – while synchrotrons have beams of different
energies for which those with energies less than around 30 keV are ideal for imaging.
Resolution is determined by the sample size, beam quality and the detector specifications; for
cone-beam set-ups (in laboratory-based instruments) resolution is also controlled by the
proximity of the sample to the beam, as mentioned previously, while detecting adsorption at
a sufficiently fine resolution. Current micro-CT scanners will produce images of around 10003-
20003 voxels. To generate a representative image, the cores are normally a few mm across,
constraining resolution to a few microns; sub-micron resolution is possible using specially
designed instruments and smaller samples. Developments in synchrotron imaging may allow
much larger images to be acquired, but at present most images have an approximately 1000-
fold range from resolution to sample size.
Pore-space images of three quarry carbonates: (a) Estaillades; (b) Ketton; (c) Mount
Gambier. The images shown in cross-section in Figures 1(a), (b) and (c) have been
binarized into pore and grain. A central 10003 (Estaillades and Ketton) or 3503 (Mount
Gambier) section has been extracted. The images show only the pore space.
Micro-porosity – small pores typically within larger grains with a size of around a micron or
smaller – can be imaged using electron microscopy techniques. The figure below shows
images of Ketton and Indiana showing small pore spaces, smaller than 1 micron (m), that are
generally below the resolution of micro-CT scanners, but which may contribute significantly
to the connectivity and porosity of the rock.
Scanning electron microscope images for Ketton limestone (top) and Indiana limestone
(bottom) at 2000x and 4000x magnification respectively, showing micro-porosity: small
pores within larger grains.
The final conceptual step is to describe the pore space of the rock in terms of a network. This
is a topologically representative description of the pore space, where the larger voids between
grains are called pores and these pores are connected together through narrower
connections, called throats. Each pore and throat in reality has a complex shape in cross-
section, but we will describe these – for simplicity – as triangles. This allows the wetting phase
to reside in the corners of the pore space while the non-wetting phase occupies the centres.
This way of viewing the rock allows us to understand multiphase flow and – in some cases –
make quantitative predictions of flow and transport properties. I will not go through the
details of how a network is extracted – there are several different methods to do this (see, for
instance, Dong and Blunt, 2009); I will simply show some illustrative examples.
The figure below shows the pore space and network for Berea sandstone, a benchmark used
for many experiments and modelling studies.
Figure showing (a) the void space of a simple sandstone (Berea) and (b) the associated
topologically representative network of pores and throats.
The next figure shows the more complex networks for carbonates – based on the images
shown previously.
Pore networks extracted from the images shown previously: (a) Estaillades; (b) Ketton;
(c) Mount Gambier. For illustrative purposes, only a section of the Mount Gambier
network is shown. The pore space is represented as a lattice of wide pores (shown as
spheres) connected by narrower throats (shown as cylinders). The size of the pore or
throat indicates the inscribed radius. The pores and throats have angular cross-sections
– normally a scalene triangle – with a ratio of area to perimeter squared derived from the
pore-space image.
8. PRIMARY DRAINAGE
Now consider a porous medium that is initially fully saturated with water, and is water-wet.
Then a non-wetting phase (oil) enters the porous medium. Imagine that this is done
sufficiently slowly that the pressure drop across the oil (from Darcy’s law, described later) is
small in comparison with the capillary pressure. This process is called primary drainage and is
the process by which oil migrates from source rock to fill a reservoir. It is also the process by
which injected carbon dioxide displaces brine in a storage aquifer.
If we return to the Young-Laplace equation (6.1), the non-wetting phase will preferentially fill
the larger pore spaces, where the radius of curvature for the meniscus is larger, resulting in a
lower capillary pressure. A lower capillary pressure means that – for a given wetting phase
pressure – a lower non-wetting phase pressure for invasion. As the non-wetting phase
pressure is increased, smaller regions of the pore space (lower radii of curvature) can be
accessed. As a consequence, primary drainage proceeds as a sequence of filling events,
accessing progressively smaller pores. In a network representation, filling pores is easy, since
they are larger than the connecting throats. Hence, the invasion of the non-wetting phase is
limited by the throat radius. The non-wetting phase will next fill the largest-radius throat that
is connected to a pore already filled with non-wetting phase. This largest throat and the
adjoining pore fill, and then again the largest-radius throat is filled. This is technically known
as an invasion percolation process (Wilkinson and Willemsen, 1983): the pore network is filled
in order of size, with the constraint that the invading (non-wetting) phase must be connected
to the inlet. There are subtleties associated with trapping the wetting phase (how does the
water escape), but this is a good model of primary drainage and motivates why a network
representation of the pore space is useful for the understanding of fluid displacement.
At a macroscopic – core – scale, if we average the behaviour over millions of individual pores
and throats, we can plot the capillary pressure (the pressure difference between the phases)
as a function of water saturation. Similar to the previous discussion of contact angle, the
process is always limited by the most difficult step, or the invasion with the highest capillary
pressure. In theory, for a slow displacement at a fixed flow rate, the capillary pressure can
decrease when larger pores, or the large throats connected to them, are filled. However, in
general, in experiments, the capillary pressure simply increased in increments, and the
amount of the pore space accessed at that pressure is recorded.
Capillary
pressure
0
0 Swc
Water saturation 1
The oil invades progressively smaller regions of the pore space. We find a connate or
irreducible water saturation Swc where further increases in capillary pressure result in little or
no decrease in water saturation. At this point the water may either be trapped in wetting
rings around rock grains, or (more likely) contained in roughness, grooves and corners of the
pore space. While, in theory, this water could be displaced, it would require a huge amount
of time and a very high capillary pressure to do so.
During and after primary drainage, regions of the pore space that come in direct contact with
oil may alter their wettability as described before (section 6.5).
In most reservoir sandstones pore radii, R, are in the range 1 – 100 m – see the previous
pore-space images. Pc 2/R (see section 6 – assuming a contact angle close to zero). ow
50 mN/m for say alkane/water. Pc is around 0.1/R or in the range 103 – 105 Pa. These values
are typically ten times higher for mercury/air, since is higher (the interfacial tension is
around 480 mN/m).
The figure on the next page shows the effect of heterogeneity and permeability on the
capillary pressure: lower permeability normally reflects smaller pore (throat) sizes and a
higher capillary pressure, while a more heterogeneous structure leads to a wider range of
capillary pressures as more of the pore space is invaded.
Capillary
pressure
Decreasing
permeability
More
heterogeneous
Relatively
homogeneous
Schematic of the primary drainage capillary pressure showing the effect of permeability
and heterogeneity in the pore size distribution.
The standard method to measure primary drainage capillary pressure is through mercury
injection. Here a small, dry rock sample – usually around 5 mm across – is placed under
vacuum and then mercury is injected as the non-wetting phase. The volume of mercury that
enters the rock sample is recorded as a function of the imposed pressure. Here we do not see
any irreducible saturation, so the mercury can, in theory, invade the entire (connected) pore
space.
It is also possible to measure capillary pressure on experiments with two fluids – such as water
and oil – although this is more difficult. Later I will show some mercury injection capillary
pressure curves measured on different rock samples. I will also present capillary pressures
measured when a core is initially fully saturated with brine and then oil (or carbon dioxide) is
injected at some known pressure and the volume that is injected is recorded. There is a
porous plate at one end – a porous ceramic disc with a very high capillary pressure – that
prevents any injected non-wetting phase leaving the system.
To help illustrate the results, overleaf I show pictures of the rock samples studied taken with
X-rays at a resolution of around 7 microns, where the pore space is clearly visible. We have
seen some of these samples previously. Then the measured capillary pressures for primary
drainage are shown: the apparatus used to make the measurements was shown above. Note
how there is a region of relatively low pressure when the non-wetting phase invades the larger
pores followed by a steep rise where smaller and smaller pores and throats are invaded. The
magnitude and shape of the curve is an indicator of pore size and structure.
Micro-CT images of four rock samples (from left to right: Indiana, Berea, Doddington,
Ketton) on which capillary pressure was measured. The table shows their properties.
Note how the trend is permeability is evident from the grain sizes: in all cases the
diameter of the sample is 5 mm.
Measured primary drainage capillary pressure: note how the magnitude of the pressure
relates to the average pore size, evident in the micro-CT images.
9. IMBIBITION
Imbibition is the opposite process to drainage, where wetting fluid invades a porous medium
containing non-wetting fluid. We normally only consider secondary imbibition, which is the
invasion of wetting fluid into non-wetting fluid after primary drainage. That is, there is some
wetting fluid initially present in the porous medium. This is the process that occurs when
water is injected to displace oil in a reservoir, when the aquifer encroaches into a gas field
during production, or when brine displaces stored carbon dioxide, as the carbon dioxide rises
in the storage aquifer.
The capillary pressure for imbibition is always lower than for primary drainage. There are
three reasons for this:
In drainage, the non-wetting fluid advances through the porous medium by a connected
piston-like advance. That is, regions can only be filled with non-wetting fluid if they are
adjacent to a region that also contains non-wetting fluid. Both wetting and non-wetting
phases remain connected. In imbibition, the displacement process is different, as described
below.
In imbibition, it is possible for the wetting phase to trap the non-wetting phase. This is through
bypassing and snap-off. Bypassing is when invading fluid surrounds and strands a ganglion of
non-wetting phase: this occurs due to local inhomogeneities in the pore structure (or local
capillary pressure). However, the more important process is snap-off. Here water flows
through wetting layers and fills narrow regions of the pore space in advance of the main
wetting front. This is the principal mechanism by which non-wetting phase is surrounded and
trapped.
For reference I show some schematic pictures of pore-scale displacement below. As the
pressure in the wetting phase increases, wetting layers in the pore space thicken. There
comes a point at which the meniscus between the wetting and non-wetting phases loses
contact with the solid – it is no longer possible to place the interface in the pore space. At this
point, the throat fills rapidly with wetting phase.
In imbibition, piston-like advance is favoured in the narrow throats, but impeded by the wide
pores (the water wants to be in the narrow regions of the pore space). However, there is a
subtlety: it is easier (that is the radius of curvature is smaller) for the wetting phase to fill a
pore if more of the surrounding throats are also full of water. This is shown in the diagram
below: in a series of classic papers Roland Lenormand (see, for instance, Lenormand et al.,
1983) described these imbibition processes as In, where the n refers to the number of
connecting throats filled with non-wetting phase. I1 is more favoured than I2 which, in turn, is
more favoured than I3. The result of this is that – at the pore scale – the wetting front tends
to be flat, filling in any local channels filled with non-wetting phase. This tends to suppress
trapping. Hence, without snap-off – described next in more detail – most of the non-wetting
phase is recovered from the porous medium (which is good for oil and gas recovery, but bad
for carbon dioxide storage).
Pore-filling processes in water (wetting phase) injection. (a) I1, when all but one of the
connected throats is full of water. This is the most favourable displacement. I2, where
two throats are initially filled with oil. This is less favoured as the threshold capillary
pressure is lower – the radius of curvature of the interface as it invades is larger.
Non-wetting
phase
Wetting phase
Non-wetting Wetting
phase phase
The snap-off process. Here water flows along the corners of the pore space indicated
in the top figure, showing a throat in cross-section. An instability occurs when the
wetting layer in the corner loses contact with the surface – any further increase in water
pressure leads to the rapid filling of the centre of the throat. The lower figure illustrates
how the non-wetting phase is pinched-off when viewed along the length of the throat.
In the snap-off process, as the wetting phase pressure increases, the wetting layers in the
corners of the pore space swell, as mentioned previously. It is possible that these layers swell
to a sufficient thickness that they lose all contact with the solid. This always occurs in the
narrowest portions of the pore space – the smallest throats. This creates an unstable
configuration, and the wetting phase then rapidly fills the throat. This process only occurs for
slow flow – there has to be sufficient time for the wetting phase to flow along wetting layers,
and for low contact angles and sharp corners (which allow the wetting layer to swell in the
first place). So, all the narrow regions of the pore space – anywhere in the rock – are filled. If
we fill all the throats around a pore, then the non-wetting phase in the pore is trapped – it
cannot escape. This is the origin of residual saturation, a very important concept in multiphase
flow that will be discussed in further detail below.
As mentioned above, snap-off is favoured for low contact angles, while connected piston-like
advance dominates as the contract angle becomes larger (but is still less than 90 o, as we only
consider water-wet systems here). Eq. (6.5) gave the capillary entry pressure for piston-like
advance for invasion of a cylindrical throat of radius r. We now consider which process is
favoured – snap-off or piston-like advance. For snap-off, we do have to have an angular pore.
If the pore is square in cross-section, it is easy to calculate the critical radius of curvature at
which the meniscus first loses contact with the solid.16 This gives a critical capillary pressure:
= (1 − )
(9.1)
where r is now the inscribed radius of the pore (or throat). The ratio of the capillary pressure
for snap-off to piston-like advance for the same throat is:
, 1
= (1 − )
, 2
(9.2)
This is always less than one: physically this is because in piston-like advance we have curvature
in two directions – the hemispherical meniscus – whereas in snap-off we only have curvature
normal to the length of the throat.
The process with the highest capillary pressure is favoured in imbibition and hence – if
possible – piston-like advance will occur rather than snap-off. However, this requires that an
adjoining pore is also full of wetting phase, which may not be the case. So, snap-off does
occur, but only when pore-filling is supressed because of the large pore size and different
cooperative pore filling mechanisms. This means that porous media with a large difference
between pore and throat sizes will see a lot of snap-off and trapping, while porous media with
similar sized pores and throats will see less trapping and a more connected wetting phase
advance. Furthermore, Eq. (9.2) indicates that snap-off becomes less favourable as the
contact angle increases. This is true even if we include different pore shapes and the effects
of cooperative pore filling: as the contact angle increases there is less snap-off and a lower
residual saturation.
16
This is a geometric calculation which can be generalized for any half-angle of the throat.
In this case the tan term in Eq. (9.1) is multiplied by tan.
During imbibition the water pressure increases, meaning that the capillary pressure decreases.
Imbibition ends at a water saturation 1 – Sor, where Sor is the residual oil saturation. This
residual oil is very important, as it determines how much oil can be recovered from a reservoir.
It is also significant in carbon dioxide storage, since this residual is trapped, cannot move and
therefore cannot escape back to the surface.
We can image residual saturations directly using micro-CT scanning: the figure overleaf shows
these trapped clusters (the non-wetting phase is dense carbon dioxide at high pressures and
temperatures) in Doddington sandstone. As mentioned above, the principal process
governing the amount of trapping is snap-off.
We can also illustrate the saturation distributions of carbon dioxide after primary drainage
and brine injection in a carbonate – Ketton, whose pore space has been shown previously.
Last, we show images from different rock samples. In all cases approximately 2/3rds of the
saturation initially present in the pore space is trapped.
These figures show that ganglia of many sizes are trapped, from clusters filling a single pore
to large clusters that almost span the system. Indeed we see an approximately power-law
distribution of cluster sizes consistent with percolation theory, which indicates that, as
expected, the pore space is filled in order of size with the smaller regions of the pore space
filled first by water, trapping the non-wetting phase in the bigger pores (Andrew et al., 2013;
2014).
A micro-CT image of Doddington sandstone showing the residual CO2. The colours
indicate the size of trapped cluster. The image has a resolution of approximately 10
microns and water and rock are not shown. The overall residual saturation is 25% (from
Iglauer et al., 2011).
The figure overleaf shows how the micro-CT images are processed to identify trapped
non-wetting phase. The analysis consists of 4 steps; filtering of the raw image (top
figures), cropping the image to the desired size, watershed seed generation to identify
different phases – the seeds are placed on voxels that are clearly one phase or another
(middle row) and the application of the watershed algorithm (bottom). The raw image
(top left) shows super-critical carbon dioxide, scCO2 (the darkest phase), brine (the
intermediate phase) and the rock grains (the lightest phase). The rock grains are around
700 µm across. The processed image shows rock as dark blue, brine in green and carbon
dioxide – the non-wetting phase – in red (from Andrew et al., 2013). Here two-
dimensional cross-sections of three-dimensional images are shown.
6 mm
3mm
3mm 3mm
Visualization of the fluids in the pore space of Ketton limestone at the end of primary
drainage. The pale blue represents a connected cluster of non-wetting phase (carbon
dioxide, CO2, as a dense, supercritical phase, at high pressure and temperatures, typical
of conditions in a deep storage aquifer). The other colours represent smaller
disconnected clusters of the carbon dioxide (Andrew et al., 2013).
6 mm
Three-dimensional rendering of CO2 after brine injection. Each unique CO2 ganglion is
displayed as a different colour. Each ganglion is isolated, and so is trapped (Andrew et
al., 2013).
Three-dimensional rendering of CO2 after brine injection. Each unique CO2 ganglion is
displayed as a different colour. Each ganglion is isolated, and so is trapped. Left
Bentheimer sandstone, middle Estaillades limestone, right Mount Gambier limestone.
The results from five experiments from each rock type are shown: the top left image
shows the fluid distribution after primary drainage, while the other five are shown after
waterflooding. The bottom row shows two-dimensional slices of the raw images. From
Andrew et al. (2014).
The final sequence of saturation change is secondary drainage, where non-wetting fluid re-
invades the porous medium after imbibition. Typical capillary pressure curves for the flooding
sequence – primary drainage, waterflooding (imbibition in this case) and secondary drainage
– are shown for a water-wet rock.
Capillary
pressure
Primary
drainage
Imbibition
Secondary
drainage
0
0 Swc 1-Sor 1
Water saturation
Why is the capillary pressure for secondary drainage lower than for primary drainage? This
has to do with the trapped non-wetting phase that becomes reconnected during secondary
drainage. Rather than think of the curve as being lower, consider the secondary drainage
shifted along the saturation axis, to represent the trapped saturation.
The key features to note in the capillary pressure are the irreducible wetting phase saturation,
the residual non-wetting phase saturation, the shapes of the curves and their relative
magnitude.
The figure below shows a schematic of different saturation paths, where the non-wetting
phase is injected to an initial saturation and then wetting phase is injected. The amount of
trapping depends on the initial saturation: as the non-wetting phase invades progressively
more of the pore space, there are more places where it can be trapped. This phenomenon is
observed in the transition zone of oil fields (discussed later) and during carbon dioxide (CO 2)
injection, where it is unlikely that the injected CO2 will completely fill the pore space
everywhere, as illustrated in the previous section.
The physical picture is as follows. During primary drainage the non-wetting phase fills
progressively smaller portions of the pore space. If primary drainage stops at some
intermediate saturation, then only the larger pores have been filled. During water flooding
(imbibition), trapping occurs preferentially in the larger pore spaces. Hence, the more of these
pores that have been filled initially with non-wetting phase, the more that can be trapped.
However, notice the characteristic curvature of the trapping curve (the relationship between
initial and residual saturation shown in the subsequent figure): at low initial saturations, only
the very largest pores are invaded and these are trapped, so the curve has a slope of almost
one (which represents everything being trapped), but the slope decreases with initial
saturation. When the initial saturation is high, only small pores are being filled and these
contribute very little to the overall amount of trapping.
A schematic of primary drainage and imbibition capillary pressure curves where primary
drainage ends at different initial saturations. This in turn determines the amount trapped
during subsequent imbibition (water flooding).
The trapping curve – the relationship between initial and residual non-wetting phase
saturation – based on the capillary pressure curves shown previously.
The next figure shows an experimentally-measured trapping curve for Berea sandstone from
the Imperial College PhD thesis of Rehab Al-Maghraby (see also Al-Maghraby and Blunt, 2013).
The upper set of points are for an oil/water system which is assumed to be strongly water-
wet with lots of trapping. Less trapping is observed when the non-wetting phase is super-
critical (sc) carbon dioxide: here it is hypothesized that the contact angles are larger,
representing a weakly water-wet system that has, consequently, less snap-off.
Thwe
Trapping curves are normally fit by empirical curves: these have no physical significance, but
are a convenient way to make sense of the data and provide convenient input into numerical
reservoir simulators. The most used model is due to Land (1968) and was originally developed
for the trapping of gas. The residual saturation is written:
∗
∗
= ∗
1+
(9.3)
where C is a constant fit to the data and S* is a normalized saturation, defined by:
∗
=
1−
(9.4)
Another model by Spiteri et al. (2008) – used in the figures above to match the data – is to
assume a quadratic (parabolic) match as follows:
= −
(9.5)
As discussed above, in general, less trapping implies less snap-off and hence a less strongly
water-wet system (larger contact angles). When the system becomes oil-wet or mixed-wet,
we see a different behaviour which is discussed later.
The Leverett J function is a way of expressing the capillary pressure in dimensionless form,
which takes account for different average pore size and interfacial tensions. This is a very
useful scaling for dealing with laboratory measurements that may be performed with fluid
pairs and at conditions – in terms of average porosity and permeability – different from in the
field.
( )= ( )
(10.1)
The J function only includes information about the geometry of the porous medium.
Sometimes the cos term is ignored, and it is assumed that the system is strongly water-wet
(cos =1) for primary drainage. For imbibition, the cos scaling is no longer appropriate, since
other displacement mechanisms (snap-off and cooperative pore filling) control the behaviour
and the contact angle term is always neglected.
Below is a measured mercury injection capillary pressure rescaled as the J-function for Berea
sandstone. Mercury is always the non-wetting phase, and so this represents a drainage
displacement. Note that the minimum value of the J function – the dimensionless entry
pressure – is typically less than 1 (around 0.2-0.3) in the example above. By a J-function value
of 1, most of the pore space has been accessed by the non-wetting phase. Much larger values
are possible, as the non-wetting phase forces its way into narrow corners and cracks of the
pore space as well as the smallest pores themselves. However, often these high values are
simply experimental artefacts: a huge capillary pressure is imposed and insufficient time is
given to achieve capillary equilibrium. In general, be cautious about using J-function values
much above 1 in quantitative calculations, as they may not correspond to a true state of
equilibrium.
100
10
J(Sw)
1
0.1
0.01
0 0.2 0.4 0.6 0.8 1
Sw
An example Leverett J-function measured by mercury injection during primary drainage
on Berea sandstone. Not that the majority of the displacement occurs for J<1. Also,
since mercury displaces a vacuum there is no irreducible or connate wetting phase
saturation in this experiment. This and the other figures in this section are taken from
the Imperial College PhD thesis of Rehab El-Maghraby.
In carbonates, however, with micro-porosity, there may be significant displacement for J>1,
as the non-wetting phase enters these small pores, as shown in the figures below.
The Leverett J-function measured by mercury injection during primary drainage, and by
displacement of brine by CO2 on Indiana limestone. Notice that when different
measurements with different fluids are represented as a J function they lie on the same
curve to within experimental error. Also note that the curve shows regions where the
saturation changes rapidly with pressure, indicting at low pressure the intergranular
porosity and, at high pressure, the micro-porosity within grains.
The Leverett J-function measured by mercury injection during primary drainage, and by
displacement of brine by CO2 on Ketton limestone. As in Indiana, there is a clear
signature of intergranular (macro) porosity and intragranular (micro) porosity.
It is possible to relate the capillary pressure to the pore size distribution: this is routinely
performed on the results of mercury injection tests. First, Eq. (6.5) is used to convert the
capillary pressure into a throat radius, assuming piston-like displacement into a circular tube.
Hence, instead of Pc as a function of saturation (the wetting phase saturation, even though for
mercury injection this is a vacuum), we define an effective radius as a function of saturation:
2
( )=
( )
(10.2)
( )= = =− =−
(ln ) ln
(10.3)
G(r) is an indication of the number of throats of radius r. Logarithmic axes are used since there
is typically a wide variation in capillary pressure and hence effective pore size. It is not strictly
the throat size distribution, since the displacement process during primary drainage –
technically similar to invasion percolation – allows the filling of regions with wide pores and
throats which are only accessed through a smaller throat at a high pressure. However, it does
give some indication of the range of pore sizes in the material.
Below I show some example distributions on our example rock types: note that for the
carbonates we see a bimodal distribution showing macro-pores and microporosity.
The mercury injection capillary pressure for Indiana limestone (bottom). This is
converted into saturation as a fraction of effective throat radius (middle) from which a
throat size distribution is computed (top) using Eqs. (10.2) and (10.3). Here we see a
very wide range of throat size with a clear indication of macro and micro-porosity.
The relationship between apparent radius and saturation, together with the inferred
throat size distribution for Ketton limestone. Here there is a clear distinction between
the large, intergranular, pores and the much smaller intragranular micro-porosity.
The throat size distribution for Doddington sandstone. Here there is a relatively narrow
distribution of large pore spaces – here there is no microporosity.
The throat size distribution for Berea sandstone. Here again there is a relatively narrow
distribution of large pore spaces, but they are smaller on average, and have a larger
range of size than for Doddington.
As we discussed before, most reservoir rocks contain both oil-wet and water-wet regions. This
means that during water invasion a negative capillary pressure (a water pressure higher than
the oil pressure) needs to be applied to force oil out of the oil-wet regions.
Below are some classic experiments (Killins et al., 1953) illustrating the effect of wettability.
The curves are good for illustrative purposes but do not show the correct residual for the oil-
wet case, since this is difficult to measure with any accuracy experimentally. I will provide a
physical explanation in this section. The emphasis is how to relate the macroscopic properties
– in this case capillary pressure – to the pore-scale physics and the configuration of fluids at
the micron scale.
The degree of trapping is dependent on the presence and connectivity of oil layers sandwiched
between water in the corners and water in the centres of oil-wet pores. This is illustrated on
the next page. If the surface is oil-wet, then water becomes the non-wetting phase. This
means that it preferentially fills the centres of the largest pores. However, water is still
retained in the corners after primary drainage. Hence, between water in the corners and
water in the centre, there is an oil layer.
These oil layers maintain connectivity of the oil phase down to low saturation. As the water
pressure increases, the oil layers become increasingly thin and will eventually become
unstable, allowing trapping. However, the flow rate through these layers is very low and so –
experimentally – you have to wait a very long time to see the oil drain down to its true residual
saturation. This situation is similar to the irreducible water saturation in primary drainage:
again we have layer flow and waiting longer can allow lower water saturations to be achieved.
Capillary pressure (Pc) curves, for water-wet, mixed-wet and oil-wet rock. Note that the
capillary pressure becomes negative if the sample is not water-wet (Killins et al., 1953).
The term “imbibition” is here incorrectly used to describe forced displacement at a
negative capillary pressure – in these notes we confine imbibition to refer only to a
spontaneous process occurring at a positive capillary pressure.
Possible configurations of oil and water in the pore space of a single pore or throat.
Initially the porous medium is completely saturated with water. After drainage, the non-
wetting phase (oil) resides in the centres of the pore space, with water confined to the
corner. The regions of the pore space directly contacted by oil may change their
wettability. When water is injected, the water can fill the entire pore space, or – if the
altered wettability surface is oil-wet – a layer of oil can form sandwiched between water
in the corner and water in the centre. This oil layers allows the oil to remain connected
and drain to very low saturation, albeit very slowly.
The figure overleaf represents a weakly water-wet system. Here some water is displaced
during forced water injection. However, the medium does not spontaneously imbibe any oil,
indicating that there are no connected oil-wet pathways through the system.
The region of the water invasion curve where the capillary pressure is negative is called forced
water injection. Where the capillary pressure is positive we have, as before spontaneous
imbibition, or spontaneous water injection. To avoid confusion, imbibition and drainage is
used only when the corresponding capillary pressure is positive.
The capillary pressure becomes negative, even though no part of the pore space has an
intrinsic contact angle greater than 90o; instead, thanks to contact angle hysteresis, or the
exact meniscus configuration during pore filling, some of the displacement pressures are
negative, meaning that a higher pressure in the water than the non-wetting phase (oil) is
required for invasion.
A mixed-wet porous medium spontaneously displaces both oil and water –– there are
continuous pathways of both water-wet and oil-wet patches in the pore space. The oil
imbibes in the same way as water in a water-wet system during secondary invasion of oil, with
snap-off of water accommodated through oil layer flow. This can lead to considerable
trapping of water as a non-wetting phase. In contrast, there is less trapping of oil, again due
to the connectivity of oil layers.
When the capillary pressure is negative during waterflooding, then the largest oil-wet pores
are filled preferentially in what is technically a drainage process, followed by progressively
smaller regions. The capillary pressure curve tends to be relatively flat as it crosses zero, since
we transition from the filling of large water-wet pores (for spontaneous imbibition) to large
oil-wet pores (for forced displacement): there is a large change in saturation associated with
the filling of these pores, with relatively little change in capillary pressure.
Capillary
pressure
Primary
Spontaneous drainage
imbibition
Secondary
drainage
0 Water
Swc 1-Sor 1
saturation
Forced water
injection
Typical capillary pressures for a weakly water-wet system. Note that there are some
parts of the pore space that require forced water injection to access: it is unusual in
natural samples to have a capillary pressure curve during waterflooding that is entirely
positive. Only the waterflooding and secondary drainage curves are shown: generally
primary drainage shows only water-wet characteristics (either mercury injection is used,
or the displacement occurs before oil has altered the wettability of the system).
Capillary
pressure
Forced oil
invasion
Spontaneous
imbibition of
water
0
Swc Water 1-Sor 1
saturation
Spontaneous
Forced water imbibition of oil
injection
Typical capillary pressures for a mixed-wet system. Here there are connected regions
of both water-wet and oil-wet pores and we see imbibition (spontaneous uptake) of both
oil and water.
Capillary
pressure
Forced oil
invasion
0
Swc Water 1-Sor 1
saturation
Spontaneous
Forced water oil invasion
injection
Typical capillary pressures for an oil-wet system. We see no water imbibition, but a
significant amount of spontaneous displacement (imbibition) of oil.
The final figure above is a strongly oil-wet system that has no spontaneous imbibition of water.
The residual oil saturation is lower that the water-wet medium; this is because the oil
maintains connectivity down to low saturation thanks to the presence of oil layers. There is
significant imbibition (spontaneous uptake) of oil during secondary oil invasion.
In a mixed-wet system the relationship between initial and residual saturation becomes more
complex than we showed before for water-wet media. First, the remaining oil saturation is
controlled by the extent to which oil is allowed to drain from the system through layers.
Hence, as more water is injected, more oil is produced and the remaining oil saturation
decreases. As mentioned before, it is very difficult to obtain a true residual, or minimum,
saturation experimentally, as the oil continues to flow, very very slowly, until low saturations
are reached.
Second, the relationship between initial and remaining oil saturation is non-monotonic. For
low initial saturations, the remaining saturation increases with initial saturation for the
obvious reason that the more oil initially present, the more oil that can be trapped. However,
at higher initial saturations, the residual decreases. This surprising phenomenon is due to the
presence and stability of oil layers. At higher initial oil saturation (and hence imposed capillary
pressure in primary drainage), the water is pushes further into the corners of the pore space.
This makes oil layers, formed during subsequent water flooding, thicker and more stable –
meaning a more negative capillary pressure is required (higher water pressure) to collapse
them. This extends the layer drainage regime, allowing more oil to be recovered. At the very
highest initial saturations, the remaining oil saturation may increase again, as there is more
oil present, and there is a subtle competition between displacement and connectivity in the
oil-wet and water-wet regions of the pore space, and the layers do, eventually, lose
connectivity.
This curious behaviour has been seen experimentally, and can be predicted using pore-scale
modelling. Some experimental data is shown overleaf.
During primary oil migration (primary drainage) into a hydrocarbon reservoir, both the
saturation and capillary pressure varies with height above the free water level – this is defined
as the depth when the oil and water pressures are the same, so the capillary pressure is zero.
Above the free water level, the capillary pressure increases with depth, as analysed earlier for
a meniscus in a capillary tube, Eq. (6.6). For a reservoir with variable permeability and
porosity, we invoke J-function scaling to determine the water saturation at a given height h
above the free water level. It is the saturation such that:
Δ ℎ= ( )= ( )
(11.1)
where is the density difference between oil and water. This can be rewritten:
Δ ℎ
(ℎ) =
(11.2)
This initial distribution of saturation – normally established from a primary drainage capillary
pressure curve is shown in the figure overleaf for a homogeneous medium. The initial
saturation, in turn, affects the wettability for subsequent waterflooding once the oil has aged
the surface: higher capillary pressure (low initial saturation) forces the oil into contact with
more of the solid, favouring oil-wet conditions; intermediate saturations may lead to a mixed-
wet rock, while a high initial water saturation means that there is little initial contact between
the oil and solid and the medium remains largely water-wet. Later we will explore the
implications for oil recovery by waterflooding, once we have discussed how multiple fluids
flow in the pore space. For now, it is sufficient to note that many oil reservoirs experience a
transition from water-wet to oil-wet conditions with height above the oil-water contact: in
low permeability rocks this transition zone may extend 10s m, as evident when representative
numbers are put into Eq. (11.1).
Imagine, for instance, a permeability of around 10 mD (10-14 m2), a porosity of 0.16, a density
difference of 200 kgm-3 and cos = 0.025 Nm-1. Then, the water saturation will decline to
close to its irreducible (connate) value for J=1 (see the previous sections). This then gives a
height of around 50 m – essentially there is a variation of saturation with height across the
entire reservoir and it is wrong to assume that the initial saturation is the irreducible value.
A schematic of the transition zone, where the saturation varies as a function of height
above the free water level (FWL), defined as where the capillary pressure is zero. The
water-oil contact (WOC, or the oil-water contact, OWC) is found from down-hole log
measurements of resistivity where there is first a noticeable presence of water (low
resistivity) in the formation. This distribution of saturation also affects the wettability
during waterflooding.
This is a quantitative and useful measure of wettability measured on core samples and first
described by Amott (for a more modern description, see Anderson, 1986, 1987). Remember
that it is difficult to measure contact angles in situ (although see Andrew et al., 2014b), and in
any event, this does not directly relate to capillary pressure, so a macroscopic description of
wettability is valuable. Start with a core at waterflood residual oil.
The wettability indices for oil and water are defined by:
Clean sand and rock will have Ao = 0 and Aw = 1. We rarely see Ao = 1. Sometimes we have
rocks where one value is low (around 0.1) and the other is 0 – little or no spontaneous
displacement, implying contact angles everywhere close to 90o. However, the most common
situation, for a reservoir sample, is mixed-wettability where both oil and water spontaneously
imbibe and neither index is zero.
Many researchers, rather depressingly, are unable to cope with two numbers to represent
wettability and so use instead the (completely useless!) Amott-Harvey index Aw - Ao. Please
never do this – the concept was introduced when it was considered that there was one
uniform contact angle in the rock, and so, by definition, one of the Amott indices had to be
zero. However, this is extremely unhelpful for mixed-wet systems: there is a significant
difference between a mixed-wet rock with an Amott-Harvey index of 0 (meaning that around
half the pores are oil-wet and the other half water-wet) and an intermediate-wet rock with
uniform contact angles close to 90o that also has an Amott-Harvey index of 0. Be sure to
distinguish between these two cases.
1. Return to our discussion of the oil/water transition region in oil reservoirs. In many
reservoirs – such as Prudhoe Bay off the North slope of Alaska – the reservoir is weakly water-
wet near the oil/water contact and becomes more oil-wet with height away from the contact.
Explain in your own words why we see this wettability trend.
2. Consider equilibrium at a line of contact for the following three situations: oil/water,
gas/oil and gas/water. Derive a relationship between the oil/water, gas/oil and gas/water
contact angles and interfacial tensions. This will be derived later in the notes when we
consider three-phase flow.
3. Consider a dome-shaped gas reservoir. The gas is trapped by a layer of shale that has
a permeability 0.1 mD and porosity 0.2. The gas is in pressure communication with the
surrounding aquifer. Estimate the height of gas in the reservoir. The gas density is 200 kg/m3
and the water density is 1000 kg/m3. The gas/water interfacial tension is 60 mN/m. (Hint:
Draw a cartoon of the reservoir. Gas can accumulate until there is a sufficient capillary
pressure for gas to enter the shale layer. Use the Leverett J function and a dimensionless
entry pressure of 0.3. The capillary pressure is the pressure difference due to density
differences.)
4. Explain why a water-wet porous medium has a higher residual oil saturation than an
oil-wet medium. Explain why a porous medium with oil/water contact angles close to 90o
gives a lower residual oil saturation than a porous medium with a contact angle close to 0.
Two porous media have an Amott-Harvey index of 0, indicating that they are neither oil-wet
nor water-wet. One medium has Aw = Ao= 0, while the other has Aw = Ao= 0.4. Which
system do you expect to have the lower residual oil saturation? Explain your answer carefully.
We now introduce flow into our analysis – in particular Darcy’s law for flow in a porous
medium and provide a pore-scale basis for the relationship between flow rate and pressure
gradient. We start with a consideration of single-phase flow (one fluid – by default water –
saturates the pore space completely, and we consider the water movement and that of
solutes dissolved in the water) before extending the analysis to multiphase flow (where oil,
water and gas may all be present).
There are two fundamental concepts that we will use to describe fluid flow. This is not
specifically flow in a porous medium, but the movement of any fluid. The first concept is
conservation of mass, written as:
∇∙ =0
(12.1)
I will not derive this equation here – which is simple to do considering flow into an arbitrary
volume and applying Green’s theorem – but I will return to similar derivations for multiphase
flow later in these notes.
The second concept is the conservation of momentum for a fluid. This is the Navier-Stokes
equation that is used to describe the flow of everything from volcanic magma, to air and
oceans. It can be written as:
+ ∙∇ = −∇ + ∇
(12.2)
where is density and v=(u,v,w) is the vector of velocity, P is pressure, and µ is the viscosity.
∇∙ =0
(12.3)
Eq. (12.3) is an expression of conservation of volume and can be derived directly (assuming
incompressible fluid) directly, in much the same way as the expression for conservation of
mass.
We also consider flows where the flow field changes slowly over time, and so we can neglect
any explicit time dependence in the Navier-Stokes equation. Furthermore, flow is very slow,
compared to, say, air flows, and in this limit the term with the velocity multiplied by itself can
be ignored (physically it means that viscous forces dominate over inertial forces – this is
discussed further below – and means that we can dismiss some of the more complex effects,
namely turbulence, which occur in other, more unconfined flows). Then we are left with the
steady-state Stokes equation:
∇ =∇
(12.4)
With modern linear solvers and fast computers, it is possible to solve these last two equations
numerically on the pore-space images we have introduced previously. We are now able to
use standard desk-top computers to solve billion-cell problems.17
The concept of slow flow can be quantified through the introduction of the Reynolds number
Re which is the ratio of inertial to viscous forces for fluid flow:
(12.5)
where (as before) is the fluid density, its viscosity, v a characteristic flow speed and L a
characteristic length. In a porous medium the fluid moves slowly through the labyrinthine
interstices between rock grains, or through a complex fracture network. The characteristic
length is that of a pore throat, or a narrow restriction between larger pore spaces which
17
We use the OpenFoam library to solve the Navier-Stokes equation: there are a number of
excellent public domain solvers, readily downloaded from the internet that can be employed.
impedes the flow. L is typically around 10–100 m (10-5–10-4 m) for consolidated rock and of
order 10–3 m for sand and gravel. The flow speed typically 10 m/day (roughly 10 –4 ms–1) or
less for groundwater movement (natural flow speeds in arid areas can be 1 m a year or
smaller) and is one or two orders of magnitude lower in oil reservoirs. For water = 103 kgm-
3 and = 10–3 kgm–1s–1, giving R ≈ 10–1–10–3. Viscous forces dominate and we have laminar
e
flow in porous media (turbulent flow generally occurs for Re ≈ 1000).
It is possible to average the Navier Stokes equation, which describes flow of a single fluid, for
slow, laminar flow past many obstacles. The flow is averaged over a representative volume
element containing several rock or sand grains. Since the Reynolds number is low, viscous
forces predominate. A pressure gradient, or force, is required for flow at a constant velocity.
It is possible to derive a linear relation between volumetric flow rate and pressure gradient,
known as Darcy’s law (Bear, 1972 – see the list of recommended books at the beginning of
these notes), which we will consider later.
First though, I will show some illustrative examples of the pore space, pressure distribution
and flow fields. The flow is relatively uniform in the more homogeneous systems, such as a
bead pack, but is confined to a few tortuous channels in the more heterogeneous media, as
seen in many of the carbonates we study.
The pore space of the three porous media shown in the previous figure: (a) bead pack;
(b) Bentheimer; (c) Portland. Then the pressure field for flow from left to right is shown,
with red representing high values and blue low values: flow goes from high to low
pressure. The final row illustrates the flow field, with the regions of highest flow
indicated. While flow is relatively uniform through the pore space of a bead pack, in the
carbonate it is confined to a few tortuous channels (from Bijeljic et al., 2011).
The pore space of the Estaillades and Mount Gambier, with the corresponding computed
pressure and flow fields. Estaillades has very tortuous flow paths indicating a poorly-
connected pore space, while Mount Gambier, despite its geologically complex structure,
is well-connected and supports a relatively uniform flow (Bijeljic and Blunt, 2013).
Normalized flow fields for: top left, Indiana limestone; top right, ME1; bottom left, ME2;
and bottom right Ketton limestone.
The local velocity in the pore space is highly variable – indeed we see, typically, eight orders
of magnitude variation in flow speed in the samples shown above. However, there is a way
to simplify this if we are only interested in the average flow.
We find, empirically, that there is a linear relationship between flow rate and pressure
gradient: this can be derived rigorously mathematically as well. This relationship is Darcy’s
law:
=− (∇ − )
(12.6)
q is not a local flow velocity, even though it has the units of a speed. It is the volume of fluid
flowing per unit area (and this area includes both solid and pore) per unit time. It is, in
essence, the sum of all the highly variable local speeds in the overall direction of flow
multiplied by the porosity. The minus sign indicates the physically obvious fact that flow goes
from high to low pressure – that is along a negative gradient in pressure. Notice that in Eq.
(12.6) I have also included the effect of gravity: g is the vector of gravitational acceleration.18
We can help motivate this, and aid the discussion of relative permeability, by quoting the
Poiseuille law: this is an expression that relates flow rate to pressure gradient in a single
circular cylindrical tube and can be derived directly from the Navier-Stokes equation:
=− (∇ − )
8
(12.7)
where Q is the volume of fluid flowing per unit time and r is the radius of the capillary. Note
the fourth power: this means that conductance is very sensitive to the size of the channel
through which the fluid flows.
This fourth power of radius – or area squared – is unlike that encountered for, say, electrical
current. The current in a wire – with a fixed potential (voltage) drop – is simply proportional
to the area of the wire. Why? There is a flux of electrons – double to area to flow and the
current doubles too. So, why a different relationship when it is fluid flow, not electrical
current? The key distinction here is that the electrons move at a speed determined by the
potential gradient and the metal in the wire – it is independent of the cross-sectional area of
the wire. For fluid flow, however, the situation is different, since we have a no-flow v=0
boundary condition on the solid surface. This means that flow speed increases away from the
solid – governed by the viscosity – and so the flow is faster in large pores than in small pores.
18
Darcy’s law is named after Henry Darcy, a French Civil Engineer, who, in 1856, published a
now-famous book “Les fontaines publiques de la ville de Dijon.” The book is available in a
2004 English translation. The main body of the work concerns the design of a network of
pipes to bring spring water into the city. As an apparent aside, and written in an appendix, is
the first statement of this famous law, based on a series of flow experiments in sand filters.
Flowing water through sand is an extremely effective way to remove bacteria (and even
viruses) from the water: the larger organisms are simply trapped in the pore space as they
(or clumped groups of them) cannot pass through the narrower pores, or are absorbed on
the huge surface area a porous medium presents. This explains how, for instance, mountain
streams run clear drinkable water, while the nearby fields are full of cow pats and sheep
droppings. And last, why is he Henry and not Henri if he was French? Well, he had an
English wife, so anglicized his first name!
The total flux, Q, is related to the average speed multiplied by the area to flow. In this case,
both terms increase with pore size.
We will meet this behaviour later for multiphase flow: how well each phase flows is
exceptionally sensitive to the size of the pores that it moves through.
Now consider that we have an array of parallel tubes a distance d apart. Then the porosity is
r2/d2 and the Darcy velocity q is Q/d2. Then we can write Eq. (12.7) as:
=− (∇ − )
8
(12.8)
=
8
(12.9)
Note that the factor /8 is typically much less than one; if we account for tortuous flow
through a less well connected pore space, the permeability is typically 1,000 times lower in
magnitude than the area of a typical pore (radius squared). We will use this concept later
when we discuss the effects of flow rate on displacement behaviour.
We can use Eq. (12.9) to estimate a typical pore size from the more easily measured
macroscopic parameters K and . We find:
(12.10)
This relationship was used for the derivation of the Leverett J-function (see section 11) and
explains the relationship between capillary pressure and its dimensionless form.
On a lighter note, below are pictures of some famous names in this subject. I have been
unable to find a picture of the most famous person specifically for this course: M C Leverett
(of J-function fame). It is interesting to note that both Darcy and Navier were born in Dijon.
George Stokes, Claude-Louis Navier and Henry Darcy. Spot the odd one out – the
scientist who was not born in Dijon.
=− −
(12.11)
Often you need to calculate the total flow Q (with dimensions of volume per unit time)
through a system of cross-sectional area A. Since q=Q/A, Darcy’s law may be written:
=− −
(12.12)
and for many linear flows the term / can be substituted by P/L, where P is a pressure
drop over a distance L.
The hydrology literature is principally concerned with the flow of water and Darcy’s law is
often seen written in terms of a hydraulic conductivity, KH defined as:
(12.13)
where g is the acceleration due to gravity and the density and viscosity are those of water. KH
has the dimensions of length/time. KH is 1 ms-1 if the volumetric flow rate of water moving
=−
(12.14)
where:
= +
(12.15)
is the sum of the pressure head P/g and the elevation head, z, where z is the vertical
coordinate (upwards). Since we will be concerned with the flow of air and oil, as well as water,
in this course, we will use Darcy's equation in the form (12.11).
The permeability K is a property of the geometry of the porous medium. Except for gas flows
at very high speeds (if we have Stokes flow at the pore scale, there is a linear relationship
between pressure gradient and flow speed), the permeability is not a function of flow rate or
the properties of the fluid, such as viscosity and density. K has the dimensions of a length
squared. Conventionally permeability is measured in units of a Darcy (D): if there is a flow of
1 cm3s–1 of a fluid of viscosity 10–3 kgm–1s–1 or 10-3 Pa s (water) through a cube of rock 1 cm
in all directions, the permeability is 1 Darcy if there is a pressure drop of 1 atmosphere across
it (1 atm ≈ 105 Pa). 1 D ≈ 10-12 m2. Although the Darcy is not an SI unit it is a convenient
measure of permeability. For consolidated rock, the mD (milli-Darcy) unit is often used. 1,000
mD = 1 D.
Although q has the units of velocity (and is often called the Darcy velocity) it is strictly speaking
not a real flow speed. The actual flow velocity in a system with porosity is q/. Remember
that q is defined as the volume of fluid passing through the soil or rock per unit area per unit
time. Imagine that we have a slab of rock 1 cm2 in cross-section with a porosity = 0.5. If q=1
cms-1, then each second 1 cm3 of fluid enters the rock. It fills the void space. If =0.5 this 1
cm3 of fluid fills 2 cm3 of rock. Hence each second the fluid encroaches a further 2 cm into the
slab. This corresponds to a flow speed of q/ or 2 cms-1. For unconsolidated sand or gravel
is approximately 0.3–0.35. For consolidated rock, deep underground, typical values are 0.1–
0.2. For fractured rock may be as low as 0.0001–0.02, while for some vuggy carbonates,
is as high as 0.4. Soils generally have higher porosities. Loamy soil typically has =0.3, while
clays have in the range 0.4–0.85.
Description Porosity %
Uniform sand, loose 46
Uniform sand, dense 34
Glacial till, very mixed-grain 20
Soft glacial clay 55
Stiff glacial clay 37
Soft very organic clay 75
Soft bentonite clay 84
It is easy to see that for most porous sedimentary rock K should be in the range of 10–10,000
mD (millidarcies) from Eq. (12.9). A porous rock or soil is only approximately modelled by a
bundle of parallel tubes, but to estimate K we may guess that r represents a typical throat size
(10–100 m), while d, the distance between throats is the grain diameter (100–1000 m),
giving K in the range 0.1–10 D. If we account for diagenesis or compaction, which shrinks and
closes some of the pore throats, leading to a tortuous confined pathway for fluid flow, then K
may be orders of magnitude lower (as seen in deep oil reservoirs), whereas for unconsolidated
gravel, the pore and grain size is much larger and permeabilities of 1000s of Darcies are
possible.
Permeability varies widely for different types of soil or rock. As can be seen in the tables
below, typical permeability values vary by ten orders of magnitude from granite (where the
fluid flows in small, poorly connected fractures) to clean gravel, with wide well-connected
pore spaces. Aquifers, consisting of sand of silt, normally have permeabilities between 1 and
1,000 D. Bedrock and oil reservoirs may have permeabilities in the range from fractions of a
mD to 1 D. Shales have permeailties measured in nD to D (10-21 to 10-18 m2).
In the four situations shown overleaf, the porous medium has a permeability of one Darcy, a
cross-sectional area of 1 cm2, and a length of 1 cm. The saturating fluid has a viscosity of 1 cp
(1 cp = 10-3 Pa.s) and a density 103 kgm-3. For the inlet and outlet pressures shown, determine
the total flow rate Q in cm3 s-1. g, the acceleration due to gravity = 9.81 ms-2. You may take 1
2 atm 1 atm
Q Q
Q
1 atm
2 atm
Q
Q
2 atm
1 atm Q
2 atm
Q
30o
1 atm
Darcy’s law describes the flow of a single species in a porous medium — typically water or oil.
Later we will return to study flow with multiple phases. However, before broaching this
difficult subject, we will first discuss transport of species dissolved in a single phase.
Imagine that some pollutant is dissolved in the water, or consider a chemical constituent
present in the oil. The pollutant (solute) not only follows the flow of the water, but
intermingles slowly with clean water because of molecular diffusion. In this section we write
down an expression for the diffusive flux.
Imagine – as in the figure below – that we have a porous medium saturated with water which
is not moving (in fact, we could do the same analysis if we considered a container just holding
water with no porous medium at all). Initially on the left-hand-side the water is salty and on
the right it is not very salty.
x
Diffusion in a porous medium. On the left – initially – there are more particles of a solute
than on the right. The arrows represent the random velocities of the particles due to
thermal motion. The dashed line is there for illustrative purposes.
The salt (solute) particles have a random motion – this is due to thermal fluctuations which
cause the particles to move constantly in random directions. After a long time this means that
the particles will be evenly distributed in the system. While the particle motion is random, on
average particles will move from left to right simply because there are more particles on the
left to begin with. An individual particle is just as likely to move to the right or the left.
This random motion of the solute is called molecular diffusion. It tends to smear out
concentration gradients.
Diffusion in a porous medium. Eventually, the average concentration will be the same
across the sample.
What is the flux? We will define a flux as the mass of solute moving per unit area per unit
time – it has units of kgm-2s-1.
=− ∇
(13.1)
J is the flux and D is the diffusion coefficient in a porous medium. C is the concentration
measured in mass per unit volume of water. The subscript refers to the solute – different
solutes have different diffusion coefficients. Note the minus sign – the flux is in the opposite
direction to the concentration gradient. This is like Darcy’s law, where the flow is in the
opposite direction to the pressure gradient.
Why is there the factor of – the porosity? This is simply a convention. Diffusion coefficients
in porous media are lower than for bulk fluids. So the porosity factor corrects for this. In fact,
the diffusion coefficient in Eq. (13.1) is still lower than in bulk – by a factor of 2 or 3 typically,
because of the tortuosity of the pore space.
What are the units of the diffusion coefficient? It has units length2/time or m2s-1 in SI units.
The value of often very small – many low molecular weight solutes have diffusion coefficients
at room temperature in the range 10-9 to 10-10 m2s-1.
If we assume that the concentration gradient is in one direction – the x direction in the two
figures above, Eq. (13.1) simplifies to:
=−
(13.2)
The last issue to discuss is how to relate this flux to situations when the water is also moving.
If the water is moving, molecular diffusion still takes place. In this case the average motion of
the water in the flow field is added to the random thermal motion of the molecules. The Darcy
velocity q is the volume of fluid per unit area per unit time. If we multiply this by the
concentration then Cq is the mass per unit area per unit time, or a flux. This is the flux of
solute ignoring diffusion. If we consider both diffusion and advection (flow) the total flux is:
= −
(13.3)
where q is given by Darcy’s law, Eq. (12.6). In three dimensions we can write:
= − ∇
(13.4)
Now we will derive an equation for the conservation of mass of a component in a porous
medium. This component is considered to dissolved in water where C is the density of species
per unit pore volume. The porosity is . Then C is the mass of per unit volume of the
soil or rock. We start by considering flow in one direction (the x direction) through a small
volume of length x along x with cross-sectional area A.
This is an important exercise, and I will expect you to be able to go through the steps yourself.
Throughout these notes we will derive conservation equations for different situations of
interest and then solve these equations. I will also expect you to be able to derive – and solve
– new equations describing new physical phenomena that you might not have met hitherto.
Area, A
x
A schematic of transport in one dimension used to derive a conservation equation.
Referring to the figure above, the mass entering the volume per unit time = AJ(x), while the
mass leaving the volume per unit time = AJ(x+x)
Mass leaving = ( )+ ∆ + (∆ )
(14.1)
Hence the difference in the mass in minus out is - − ∆ . Now consider conservation of
mass: Mass in per unit time – mass out per unit time = rate of change of mass in the volume
Ax:
− ∆ = ∆
(14.2)
+ =0
(14.3)
We can repeat this analysis for flow along the y and z axes. If qx, qy and qz represent the flows
in the x, y and z directions respectively, the three dimensional version of Eq. (14.3) is:
+ + + =0
(14.4)
The three-dimensional conservation equation can also be more elegantly and simply derived
using vector calculus and Green’s theorem. Consider an arbitrary volume V of the porous
medium bounded by a surface S with a flux J through it. Then the equation of mass
conservation is:
+ ∙ =0
(14.5)
We use Green’s theorem to convert the surface integral into a volume integral:
+ ∇∙ =0
(14.6)
and if this is true for an arbitrary volume the integrands must be related by:
+∇∙ =0
(14.7)
We can instead perform an overall volume balance. If we assume that the fluid is
incompressible – that is the water density w is a constant – then the volume flowing into any
arbitrary volume is the same as the volume flowing out. Form the definition of the Darcy
velocity, this leads to:
∇∙ =0
(14.8)
We can then substitute Fick’s law of diffusion and Darcy’s law to find J, Eq. (13.4). If we
assume that the diffusion constant is fixed then Eq. (14.7) becomes:
+q∙∇ = ∇
(14.9)
where I have now dropped the superscript for convenience. In one dimension:
+q =
(14.10)
Eqs. (14.9) or (14.10) are called either the convection-diffusion equation or the advection-
diffusion equation. The dissolved contaminant follows the overall flow of the fluid (water) by
Darcy’s law (convection or advection) as well as diffusing through the system by Fick’s law.
Later on we will extend this equation to multiphase flow.
(14.11)
It is possible to estimate the magnitude of the terms in Eq. (14.10) and so see whether
diffusion or advection dominate. Imagine that the contaminant has spread over some distance
x = X in a time t = T. We’re not going to attempt to solve any of these equations exactly – this
comes later – but as a crude estimate we may estimate that the three terms in Eq. (14.10)
have magnitude:
+q =
+ ~
(14.12)
This is a common way of analyzing complex equations. We are going to compute the relative
magnitude of each of the three quantities in Eq. (14.12). If advection dominates, we equate
the first two terms to find:
(14.13)
If diffusion dominates, we equate the first and third terms of Eq. (14.12) to obtain:
X~√
(14.14)
Thus for advective flow the pollutant spreads linearly with time, whereas if diffusion
dominates, the spread is proportional to the square root of time. The ratio of the diffusive to
advective terms is:
Diffusion
~ = Pe
Advection
(14.15)
This defines a dimensionless Peclet number, Pe. The length scale X is generally considered to
be a representative scale in the porous medium, which is a mean grain (or pore) size: say
around 100 m (0.1 mm). Typical flow speeds q/are of the order 1 m/day or around 10–5
ms–1 and lower. The diffusion coefficient in water for most petroleum components (typical
pollutants in groundwater) is around 10–9 m2s–1. This gives a Peclet number of around 1 –
diffusion and advection are typically similar in magnitude at the pore scale.
Molecular diffusion is more significant over lengths, X less than approximately 0.1 mm
(corresponding to a time of around 10 s), but that for large flows over hundreds of metres
taking place over months and years, advection is by far the more important process.
We have derived differential equations that describe the flow of a dissolved contaminant in
an incompressible fluid. These equations can either be solved analytically for simple cases, or
numerically. Except for small scale phenomena, advection, or flow computed using Darcy’s
law, is much more significant than molecular diffusion.
Also – rather interestingly – the same conservation equation pertains within a pore, if we
ignore the porosity in the equations: we simply invoke conservation of mass in a volume of
flowing fluid. We can also apply Fick’s law (again without the porosity term), but the flow field
is governed not by Darcy’s law, but form a solution of the Navier-Stokes equation, presented
in section 12. Primitively, the movement of a dissolved solute within the pore space is
governed by the equation:
+ v. ∇ = ∇
(14.16)
where v is the local flow velocity and D is, strictly, the molecular diffusion coefficient. v is given
by the solution of the Navier-Stokes equation; it is this equation that can be used to describe
transport at the pore scale.
Strictly speaking, Eqs. (14.9) and (14.10) are upscaled versions of Eq. (14.16) where now the
Darcy velocity, q, is given by Darcy’s law. The problem – as we discuss later – is the diffusive
flux. This contains contributions not only from diffusion, but from the random motion of the
particles in a spatially heterogeneous flow field, which we have – so far – ignored.
We will now present some solutions to Eq. (14.10). These solutions also apply to heat
transport and can be found in the classic work by Carslaw and Jaeger (1946). In the end, there
is no one correct way to arrive at a solution. Instead, we can use physical inference to find a
functional form that is likely to work. Below I show the solution to the equation for solute
that is originally injected as a point source. Physically, we expect the plume to move with
some average velocity v=q/ and then spread out dependent on the degree of diffusion. This
is what we see. We will use this insight to develop a possible mathematical form of the
solution.
A schematic of the solution of the advection-diffusion equation for an initial point source
of solute at x=0. The different Peclet numbers correspond to different average velocities.
We can transform the governing partial differential equation into an ordinary differential
equation using the following variable:
−
=
√
(14.17)
where v=q/ and we assume that the solution C can then be written as follows:
( )
=
√
(14.18)
and we solve for the (unknown) function g(z). There is – in principle – no reason why this has
to be the case. All I need to demonstrate is that the solution obeys the governing partial
differential equation and the boundary conditions. Near the end of these notes, I will present
another solution for a non-linear diffusion problem that uses a different set of variables,
because the boundary conditions are different.
= − −
√ 2
(14.19)
1
=
√
(14.20)
1
=
(14.21)
+ +2 =0
(14.22)
2 + =0
(14.23)
2 + =
(14.24)
We also know that – from our schematic solution – that the concentration will be zero at large
distances from the origin (large z): hence both g and dg/dz tend to zero for infinite z. This
means that c in Eq. (14.24) is zero and we can readily integrate:
4 ln = − +c
(14.25)
( , )= ( ) ⁄
√4
(14.26)
where M is the initial mass (per unit area) of concentration. These are the solutions shown in
the figure shown previously.
This relationship, Eq. (14.26) makes use of the following identity too find the constant of
integration:
( , ) =
(14.27)
since:
=√
(14.28)
Note that Eq. (14.26) shows a mean (maximum) concentration that moves a distance x=vt,
with a typical spread x-vt = 2√(Dt), similar to our simple scaling analysis.
The governing partial differential equation and its solution are classical in the literature, as
mentioned previously. However, they are a very poor approximation of what really happens
in a porous medium, particularly the heterogeneous pore spaces we have shown earlier. The
limitation in the derivation is that we assume a uniform flow field where the only flux
associated with changes in concentration is due to molecular diffusion.
In reality – in porous media with tortuous pore spaces – the spreading and mixing of a solute
is controlled by two factors: molecular diffusion that leads to a local – pore-scale – mixing of
concentration; and variations in flow speed that allow the solute to follow different flow paths
through the system. It is this second effect that normally dominates, with huge spreading of
a dissolved plume of solute caused not so much by local-scale mixing, but governed by the
variations in the flow field.
This is evident from an examination of the flow paths shown when we introduced Darcy’s law.
In particular, refer back to the images and flow fields shown in section 12. To have an idea of
what transport really looks like in heterogeneous porous media, the figure below shows
simulated concentration profiles – for an effective point source injection as described above
– compared to measurements of water movement at the scale of a few mm to cm made using
NMR techniques. For simplicity, the curves are shown in a dimensionless form. The y axis
represents concentration times average distance moved, while the x axis is the distance
divided by average speed times a dimensionless distance. A system with no diffusion (or
dispersion) would then travel at unit speed on the graph. With simple diffusion – obeying Eq.
(14.26) – we would see a Gaussian-type profile centred on 1. This is seen for the bead pack.
Here the flow field is uniform and there is Fickian-type smearing controlled by molecular
diffusion and small-scale heterogeneity in the flow field.
For the more heterogeneous porous media – the sandstone and carbonate – the behaviour is
different: most of the solute resides in locally stagnant regions of the pore space and hardly
move, with a very long dispersed plume of solute in the faster flowing regions. There is no
obvious concept of a typical speed with smearing about this average.
A full understanding of this phenomenon and how, properly, to describe dispersive transport,
is a rich topic of current research. An exploration of the ideas and how to describe transport
mathematically in these cases is beyond the scope of these notes: needless to say an approach
based on fluctuations in velocity and/or travel times is necessary, which does not fit neatly
into an effective partial differential equation – it is evident from the figures, that we cannot
match the experiments simply by tweaking the effective diffusion (or dispersion) coefficient
while leaving the functional form of the behaviour the same.
Predicted (solid lines) and measured (dotted lines) dimensionless concentration profiles
for transport in a bead pack, sandstone and carbonate. Note that with the exception of
the beadpack, the behaviour does not resemble a solution to the governing advection-
diffusion equation. From Bijeljic et al. (2011).
If – and in reality this is rarely, if ever, the case – we can indeed describe transport as the
solution to an advection-diffusion-type equation in a heterogeneous medium, then
traditionally we can write the following identical in functional form to Eq. (14.9):
+q =
(14.29)
but where now D is the effective dispersion coefficient that accounts for both molecular
diffusion and the random nature of the flow field. This equation is called the advection-
dispersion equation. This coefficient D can be written as follows:
= + = +
(14.30)
where now Dm is the molecular diffusion coefficient in the porous medium, and is the
dispersivity – it has the dimensions of a length and represents, physically, the typical scale of
heterogeneity in the porous medium. However, natural geological media have structure over
all length scales, and so the apparent dispersivity – and hence the spreading of a contaminant
plume – appears to become more significant with scale, as more heterogeneity is
encountered. The table below illustrates this phenomenon, where the dispersivity is,
approximately, around one tenth of the scale of the system.
Eq. (14.30) can be derived assuming that the solute experiences a series of random
perturbations every time it moves a distance – the velocity v in the dispersion coefficient
represents the fact that the number of perturbations encountered in a given time is
proportional to the flow speed.
In the plane of the aquifer (a nearly horizontal plane) the plume is therefore highly dispersed
due to variations in permeability that cause huge fluctuations in the local flow rate. The
permeability K will typically vary by several orders of magnitude or more over lengths of a few
metres in heterogeneous formations. A contaminant plume does not flow at a constant rate
and direction, but forms a ragged front due to variations in permeability. This causes the
plume to spread. Small amounts of molecular diffusion mix the contaminant with clean water.
The combined effects of permeability variation and diffusion dilute the plume and disperse
the contamination over a wide region of the aquifer. In virtually all circumstances the precise
location of the contaminant cannot be predicted with any certainty unless the distribution of
permeability in the subsurface is known: at every scale from the pore-scale onwards, the
distribution of contaminant is rarely, if ever, accurately predicted by an average displacement
and some Gaussian-type variation about this mean.
Overall, the characterization of dispersion as a diffusive process is flawed, since this does not
– at any scale – the pore scale or the field scale – characterize transport in even a qualitative
sense. The community does not – at present – have a good way to describe transport on all
scales, although much of the mathematical and physical insight has been developed.
In some sense, multiphase flow, which we will now return to, is easier and we will revisit
diffusive processes in this context – specifically capillary-controlled displacement – later.
Now we will return to multiphase flow in porous media and the concept of capillary pressure.
In these notes, I presume that the fluid configurations are controlled by the Young-Laplace
equation and contact angles. Fluid flow is slow and – as we show later – we presume that
each phase flows independently.19
However, do capillary forces really dominate at the pore scale and what are the effects of
buoyancy forces and flow rate?
If a representative pore radius is R, then a typical capillary pressure is of order /R. The viscous
pressure drop over a length L is given from Darcy’s law (where P is a pressure drop – hence
the removal of the minus sign):
∆
q= ;∆ =
(15.1)
Ratio =
(15.2)
LR/K is a dimensionless ratio (consider the units – permeability has the units of length
squared). In porous media this ratio is typically around 1,000 if L is a pore length. Typically,
L/R is around 2-10.
Why isn’t this ratio LR/K closer to 1? Recall the discussion in section 12, which showed that
permeability typically has a numerical value that is much lower than the square of a typical
pore radius.
19 This approximation can be relaxed and dynamic models of multiphase flow show that there
is viscous coupling between the phases. However, in these notes, we will ignore these
effects.
(15.3)
Representative values for the capillary number for field-scale displacement are typically 10 -8
to 10-6 or lower. If Ncap is around 0.001, then viscous and capillary forces are approximately
equivalent at the pore scale – see Eq. (15.2).
In most natural flows in aquifers and oil reservoirs, capillary forces dominate at the pore scale:
q is generally around 10-8 to 10-5, is around 10-3 Pa.s (for water), while for oil/water systems
is typically 0.05 N/m at ambient conditions and around half that number at oil field
temperatures. This leads to values of capillary number in the range 10-6 and lower. Viscous
forces, however, dominate at the large (inter-well) scale: this can be seen by substituting
L=100-1,000 m (the inter-well scale) in Eq. (15.2).
We can perform a similar analysis for buoyancy. The pressure drop over a vertical distance L
is gL, where is the density difference between the phases. The ratio of buoyancy to
capillary forces is given by:
∆
Ratio =
(15.4)
∆
=
(15.5)
Once again using representative numbers – say a density difference of 200 kg.m -3 and a pore
length of 10-4 m – the Bond number is of order 10-3. Buoyancy forces are small in comparison
with capillary forces at the pore scale.
The reason why this is important, is that if viscous or buoyancy forces were to dominate at
the pore scale then two things happen:
1. The non-wetting phase (oil) is rarely trapped, since the invasion of water occurs
through a flat connected front with essentially no snap-off and little bypassing.
2. Even if some ganglia of oil are stranded, viscous forces can push these blobs out of
the pore space.
The result is that if the capillary number is around 0.001 or larger, the residual oil saturation
can be very low. Therefore, if we could increase the capillary number in this range – by
increasing the flow rate q, or reducing the interfacial tension , we would have a very efficient
oil recovery process. This is physics behind surfactant flooding. Here surfactants are added
to the injected water, lowering the oil/water interfacial tension to 0.1 mN/m or smaller. If the
capillary number increases to the range of 0.001 or above, then very high oil recoveries are
observed.
The decrease in residual saturation as a function of capillary number is shown in the graph
below (taken from Lake, 1989).
We will now extend Darcy’s law to cases when multiple fluid phases are flowing. We assume
that each phase flows in its own sub-network of the pore space without affecting the flow of
the other phases. This is applicable for flow at low capillary and Bond numbers, where
capillary forces dominate at the pore scale.
For single-phase flow in one dimension, we have Eq. (12.11). Then the extension to
multiphase flow of fluid p is (first proposed by Muskat and Meres in 1933; see Muskat’s classic
textbook from 1949):
=− −
(16.1)
krp is the relative permeability of phase p. It represents the mobility of the phase as a fraction
of what it would be for single-phase flow. It is traditionally plotted as a function of saturation.
In these notes, we will accept this characterization of multiphase flow. For slow flow,
dominated by capillary forces, this is a reasonable approximation. However, the relative
permeabilities are not simply unique functions of saturation. As we found for capillary
pressure, the relative permeabilities will also depend on wettability (and this can change
during the course of a displacement, or series of displacements) and saturation history.
Moreover, at higher flow rates, the relative permeabilities will also be functions of this flow
rate, as well as viscosity ratio. Last, the flow of one phase can also affect the flow of the other,
through viscous coupling at the fluid interfaces. All of these effects may be significant,
particularly when capillary forces are no longer dominant at the pore scale. However, in this
treatment we will only consider the impact of pore structure, wettability and saturation path:
this alone reveals a rich and important behaviour which is sufficient to explain and explore
most displacement processes seen in oil fields and aquifers.
Below are typical relative permeability curves for a water-wet medium. The curves are shown
for waterflooding from the connate or irreducible water saturation.
1
kromax
kro
kr
krw krwmax
0
0 Swc Sw 1-Sor 1
Typical relative permeability curves for waterflooding a water-wet medium. The points
to note are the steep decrease in oil relative permeability with water saturation, the low
water relative permeability (including its final value) and the high value of residual oil
saturation Sor.
1. Typical values of the maximum oil and water relative permeability. At the beginning
of water injection – marking the end of primary drainage – oil fills most of the pore
space. Since wettability is altered only after oil invasion, the oil will reside in the larger
regions of the pore space, confining water to the corners and the smallest pores. As
a consequence, the oil relative permeability at the beginning of waterflooding – the
maximum value – is close to 1: typically 0.8 or greater; the water relative permeability
is zero, or close to zero, at the start of the displacement. When water is injected – for
a water-wet system – it preferentially fills the narrower regions of the pore space,
trapping oil in the larger pores. This means that the water always has poor
connectivity and the relative permeability remains low: usually at the end of
waterflooding the water relative permeability is only around 0.1 or lower. This is
different – as described below – if the system is not strongly water-wet.
2. The residual oil saturation is large. This is related to the discussion for the previous
point. Remember, in a water-wet system, water remains in the small pores, while oil
is in the large pores. The oil can be trapped in these larger regions of the pore space
by snap-off, leading to a large immobile or residual saturation at the end of
waterflooding: typically Sor is in the range 0.2-0.5 for a water-wet system.
3. Why the sum of the relative permeabilities is less than 1 for all saturations, and why
the sum is much less than when the curves cross. Any interface between oil and
water, across a pore space, prevents flow. Hence, the more interfaces between the
phases, the more flow is restricted. This is particularly true when water invades by
snap-off, cutting off oil flow through the largest pores, while remaining poorly
connected itself. Hence, two phases in combination have a much lower conductance
than for single-phase flow (one or more of the relative permeabilities is typically
always very low). When the relative permeabilities cross is usually when there is most
phase interference and both values are likely to be small.
In the following, we will amplify these points through experimental and modelling results.
Relative permeability is very difficult to measure accurately; sophisticated apparatus is used
to measure this important quantity – a discussion of experimental techniques if, however,
beyond the scope of this course. By default, the data refers to water injection, since
commonly this is the most important process involving multiphase flow in oil reservoirs. The
picture below shows some of our apparatus used to measure relative permeability at Imperial
College. There is an adapted medical X-ray scanner (with a resolution of around 1 mm) that
can monitor fluid movement within rock cores several cm across and up to 1 m long.
The key controls on relative permeability are wettability and the connectivity of the pore
structure, again discussed in more detail later in this section.
The points above can be illustrated in the relative permeability curves shown below. The
experimental results are part of a classic series of measurements made on Berea sandstone
for both two and three-phase (oil, water and gas flowing – discussed later) by Oak (1990).
These measurements have served as a benchmark for analysis in the literature, because Berea
is a standard quarry sandstone used by many researchers and the raw data is available in
spreadsheet form. The results of three sets of experiments are shown.
Measured (points) and predicted (lines) relative permeability curves in a water-wet Berea
sandstone. The curves are shown on both linear and logarithmic axes. The primary
drainage curves are shown at the top, and the waterflooding curves at the bottom.
Network modelling – capturing the connectivity of the pore space using a random lattice
of pores and throats, combined with an accurate assessment of pore-scale displacement
processes – can predict the behaviour accurately. From Valvatne and Blunt (2004).
Also shown are network modelling predictions performed by Valvatne and Blunt (2004).
These predictions use the network for Berea presented earlier in these notes and employ the
pore-scale displacement processes we have described. Primary drainage is an invasion
percolation process: it is assumed that the wetting phase (water) is strongly wetting and the
contact angle is zero. We make good predictions of the measured data. For waterflooding,
there is an uncertainty. As discussed in section 6, the (advancing) contact angle is larger in
this case, mainly due to the roughness of the solid surface. In network modelling we assign
an effective contact angle that accounts for roughness and the converging/diverging nature
of the pores: this angle is around 60o. With this larger contact angle – which tends to suppress
snap-off in some pores – we predict the relative permeability curves accurately.
The measured data and the predictions show the features mentioned above. Note that the
residual saturation is around 0.3 and the maximum water saturation is only about 0.1.
Blocking the flow in the largest 30% of the pore space reduces the water conductivity by a
factor of 10. This is an important observation: small changes in fluid configuration that
prevent flow through a few large channels have a big impact on relative permeability.
We can also study the effect of different displacement paths on the relative permeability. The
data in the curves below are taken from Akbarabadi and Piri (2013). Here carbon dioxide is
the non-wetting phase and is injected into a Berea core to different initial saturations. Then
brine is injected to displace the carbon dioxide, resulting in different curves (relative
permeability hysteresis) and different amounts of trapped non-wetting phase, as discussed
previously.
Measured relative permeability on Berea sandstone. Here carbon dioxide is the non-
wetting phase and is injected to different initial saturations before brine injection. Here
we see the relative permeability hysteresis: notice the different residual saturations,
dependent on the initial saturation, as discussed previously. From from Akbarabadi and
Piri (2013).
We can continue our study of relative permeability with more results where carbon dioxide is
the non-wetting phase (the application here is carbon dioxide storage in aquifers). The graphs
below show primary drainage capillary pressure (section 8), relative permeability and trapping
curves (section 9) on Berea sandstone, comparing the results of Krevor et al. (2012) with those
of other researchers. This is the combination of multiphase properties that control fluid
movement and recovery in the subsurface.
We can also consider the effects of wettability on the relative permeability curves. This will
be presented in more detail later in the context of carbonate rocks. In general carbonates
tend to experience a stronger wettability alteration in contact with crude oil than sandstones.
So, while we see the whole range of behaviour from water-wet to strongly oil-wet in
sandstones, more typically in carbonates we see mixed-wet to oil-wet properties. The figure
below shows predicted waterflood relative permeabilities for a mixed-wet reservoir
sandstone compared to the data. The predictions use different models to assign wettability:
there is some discussion in the literature over whether large or small pores are more likely to
undergo a significant wettability change. The results are also taken from Valvatne and Blunt
(2004).
In a mixed-wet system the key features are a low residual oil saturation (noted previously for
capillary pressure, section 10), and low oil and water relative permeabilities. The low residual
is due to the connectivity and slow drainage of oil layers. The low oil relative permeability is
also easy to explain: where the system is oil-wet, the oil resides in the smallest pore spaces
and in layers that, while interconnected, have a very low conductance. The water relative
permeability is also low: this is a significant feature that has a major impact on waterflood oil
recovery at the field scale and is discussed further later. When water is first injected, it fills,
preferentially, the water-wet regions – the smallest water-wet pores and throats. The water
saturation increases, but the connectivity of the water is poor and so the relative permeability
remains low. Then water fills the oil-wet regions – and the largest oil-wet pores first. Again,
to begin with, the connectivity remains low; it is only at the highest water saturations that
water becomes well connected through the larger regions of the pore space and the relative
permeability rises rapidly. Its maximum value is higher than for water-wet systems, since the
residual oil saturation is lower (there is more water in the rock) and the water occupies,
preferentially, the larger oil-wet pores.
Predicted and measured relative permeability for a mixed-wet reservoir sandstone. The
predictions (lines) use different assumptions as to which pores become oil-wet after
primary drainage: the results are relatively insensitive to this assignment. From Valvatne
and Blunt (2004).
The final set of sandstone curves are for an oil-wet reservoir rock. Again good predictions can
be made. In the figure below, the layer drainage regime is evident; this is where the oil relative
permeability is low, but allows flow down to a very low residual saturation. The water relative
permeability can reach high values, once water is well connected through the pore space in
the larger pores.
Predicted and measured primary drainage capillary pressure and waterflood relative
permeability for an oil-wet reservoir sandstone. Good predictions are made. Note the
low oil relative permeability at higher water saturations – this is the oil layer drainage
regime, where oil can be displaced to a very low residual saturation (less than 10%).
Once the water becomes well connected through the pore space occupying the larger
pores, its relative permeability rises quickly. From Valvatne and Blunt (2004).
There are two distinct recovery processes in oil fields when water is injected to displace brine.
The first is direct displacement, shown in the diagram on the next page: water is injected and
essentially pushes out oil. At the pore scale, we know that capillary forces dominate and the
recovery is controlled by the residual saturation. How fast this recovery occurs is controlled
by the relative permeabilities: the ideal is a low water relative permeability that holds water
back and a high oil relative permeability, allowing oil to flow rapidly and be displaced ahead
of the water. This is discussed later through a rigorous mathematical treatment of the
governing flow equations, but qualitatively, a low residual saturation says how much oil can
be recovered in theory (a low residual indicates high recovery) while the relative permeability
gives the rate at which recovery occurs.
The second process is imbibition. This is simple to imagine – it is the same as placing a piece
of rock in water: the water spontaneously enters the rock under the influence of capillary
pressure. This is the dominant recovery process in fractured reservoirs – typically seen for
brittle rocks, such as carbonates. Here the injected water, rather than forcing out the oil,
flows rapidly along the high permeability fractures. Then water enters the matrix (normal
unfractured rock) by imbibition. This is a process controlled by capillary forces, with some
help due to the density difference between water and oil that helps push the water into the
bottom of a matrix block. Again this process is shown schematically below. In this case the
recovery is controlled by how much oil remains after spontaneous imbibition: from our
discussion of capillary pressure, this saturation (when the capillary pressure is zero) is much
lower than the residual saturation for mixed-wet systems. The rate of recovery is controlled
by the water relative permeability, typically at low saturation, since this is the saturation range
of interest and limits how fast the oil can be displaced.
Both of these problems can be analysed through analytical solutions (in one dimension) to the
governing flow equations; these are presented later in these notes (sections 19 and 20). Here
we explain the results physically in terms of the relative permeabilities and capillary pressures.
Unsteady State
A schematic of the two types of recovery process in reservoirs. The top picture shows
displacement, where water is injected to displace oil. This occurs in most unfractured
reservoirs. However, where there is extensive fracturing, and these fractures provide
high permeability paths for flow, the behaviour is different: shown in the bottom picture.
Here recovery occurs by imbibition of water into the matrix (normal unfractured rock).
In imbibition, the recovery as a function of time has a characteristic behaviour that has been
studied by many authors. Shown below is a compilation of 48 datasets in the literature,
compiled by Schmid and Geiger (2012) in a classic paper that also presents a closed-form
analytic solution to the flow equations for this problem (which is presented later in these
notes).
In the figure above, the recovery is plotted as a function of dimensionless time. A full
discussion giving an analytical solution is provided in section 20; however, we can use physical
principles to estimate likely time-scales. We will show that this is a diffusive problem
mathematically, so we can readily examine the likely scaling of the displacement.
The driving force is capillary pressure, which is the interfacial tension divided by a typical pore
radius. As before (section 10) we can relate this to the square root of the permeability divided
by the porosity. Imagine that the wetting phase has invaded a distance x into the porous
medium. Hence the pressure gradient driving flow can be written:
(16.2)
Then from the multiphase Darcy law, Eq. (16.1), assuming that flow is limited by the water
relative permeability, we can find the flow rate, which determines how fast the distance x
changes with time:
1
= ~
(16.3)
The porosity term for dx/dt converts a Darcy velocity into a speed. Eq. (16.3) has the solution:
( )=√
(16.4)
where:
=
2
(16.5)
Note that the distance travelled (and hence recovery) scales – at early time, before the
imbibing front reaches the ends or boundaries of the system – as the square root of time. This
mathematically and physically is a diffusive process, as opposed to recovery by direct
displacement where recovery and front movement increases linearly with time.
Eventually, the wetting front reaches the end of the system (say a distance x=L); from then on
recovery is much slower. Empirically – simply a match to the compilation of recovery results
shown previously – we find that the recovery can be written as:
= (1 − )
(16.6)
where R is the oil recovery, R∞ is the ultimate recovery. is a constant used to match the
data and tD is a dimensionless time. This is the analytical match shown on the previous graph.
In our analysis, it would be given by:
(16.7)
However, this is not necessarily accurate, as this was a simplistic analysis; a more complex but
analytically correct expression is found in Schmid and Geiger (2012) that accounts for the flow
of both water and oil: this is presented later in section 20.
We can use Eq. (16.7) though to estimate timescales for imbibition recovery. What is a typical
imbibition time for a water-wet rock of size 1 cm? This is the real time necessary for have tD
around 1 in Eq. (16.7). Using = 0.04 N/m, w = 10-3 Pa.s, K = 10-14 m2 (10 mD) and = 0.2, we
find, for a typical end-point relative permeability value of 0.1 (a water-wet rock), times around
100s: imbibition is typically quite quick for small systems. What about a matrix block 10 m
across? Notice that the time-scale increases as length squared, and so in this case imbibition
takes a million times longer, or around 3 years.
In contrast, the waterflood recovery improves as the system becomes more mixed-wet. This
is a consequence of the lowered residual oil saturation. Also, as discussed in more detail later,
the low water relative permeability holds back the injected water, allowing oil to escape and
providing – in this case – a favourable displacement efficiency.
Waterflood recovery (left) and imbibition recovery (right) for sandstone cores aged for
the length of time (in hours) indicated on the graphs. The more the core is aged in crude
oil (essentially soaked in crude for different amounts of time), the more mixed-wet in
character it becomes. No ageing is least favourable for waterflooding, because of the
high residual oil saturation in this water-wet case, but is favourable and fastest if
recovery is controlled by imbibition (from Behbahani et al., 2005).
In this section we will go through a network analysis of relative permeability, to show how we
predict multiphase flow properties, their behaviour and how this relates to field-scale
recovery. We will also compare the results against experimental data in the literature. The
emphasis in this section will be on mixed-wet systems, which comprise the vast majority of
carbonate rocks, which in turn contain most of the world’s remaining reserves of conventional
oil, mainly in the Middle East. Most of the analysis is taken from Gharbi and Blunt (2012).
We start by showing images of the carbonates that we will study and the networks extracted
from these images. This forms the basis of the modelling.
Pore networks extracted from the images shown in the previous figure. The pore space
is represented by a lattice of pores (represented by spheres) and throats (represented
by cylinders): in cross-section each pore and throat is a scalene triangle.
A detailed description of the extracted networks is provided in the table below. The samples
cover a wide range of average coordination numbers: ME1 and Portland are poorly connected
with coordination numbers of approximately 2.5 whereas Guiting and Mount Gambier are
highly connected with average coordination numbers of 5.1 and 7.4 respectively. As we show
later, the average coordination number (average number of throats connected to a single
pore) is a key determinant of relative permeability and residual saturation. It is derived from
the network extraction analysis and is an indicator of the connectivity of the void space.
The pore and throat distributions of the networks are presented below.
Pore inscribed radius distributions for (a) Middle Eastern sample 1 (b) Portland
limestone (c) Indiana limestone (d) Middle Eastern sample 2 (d) Guiting
carbonate and (f) Mount Gambier limestone. In this and subsequent figures,
samples are presented in order of increasing coordination number: from a low
connectivity sample (a) to a very high connectivity sample (f).
Throat inscribed radius distributions for (a) Middle Eastern sample 1. (b)
Portland limestone. (c) Indiana limestone. (d) Middle Eastern sample 2. (d)
Guiting carbonate. (f) Mount Gambier limestone. Samples are presented in order
of increasing coordination number.
Capillary controlled displacement is simulated using the pore network model developed by
Valvatne and Blunt (2004). Initially the medium is assumed to be filled with the wetting phase
(brine) and oil is then injected. After oil invasion, we alter the wettability of the pore spaces
in direct contact with oil to represent mixed-wet conditions. Waterflooding is then simulated
and relative permeability curves are generated.
We study the impact of wettability in mixed-wet media where some fraction, f, of the pore
space occupied by oil is made oil-wet and a fraction 1-f remains water-wet. We vary the oil-
wet fraction from zero (a strongly water-wet case) to 1 (strongly oil-wet rock). In addition to
modelling mixed-wet media, this methodology reproduces wettability alteration which is due
to asphaltene deposition/ precipitation in carbonates. This alteration, governed by oil
composition, brine salinity and rock mineralogy is difficult to predict a priori.
Where oil has been in contact with the carbonate surface (pores and throats), random contact
angles with no spatial correlation are assigned with different distributions – given in the table
below – for the water-wet and oil-wet pores and throats.
The three-dimensional networks are composed of individual elements (pores and throats)
with circular, triangular or square cross-sectional shapes. Using square or triangular-shaped
networks elements allows for the explicit modelling of wetting layers where non-wetting
phase occupies the centre of the element and wetting phase remains in the corners. The pore
space in carbonates is highly irregular with water remaining in the grooves and crevices after
primary oil flooding due to capillary forces. The wetting layers might not be more than a few
microns in thickness, with little effect on the overall saturation or flow. Their contribution to
wetting phase connectivity is, however, of vital importance, ensuring low residual wetting
phase saturation by preventing trapping. Wetting layers of water are always present in the
corners, while layers of oil-sandwiched between water in the corners and water in the centre
can be observed in oil-wet regions. Layer drainage is when oil flows in these layers, allowing,
slowly, very low saturations to be reached.
Input parameters
Initial contact angle (degrees) 0
Interfacial tension (mN/m) 48.3
Water-wet contact angles (degrees) 0-60
Oil-wet contact angles (degrees) 100-160
Oil viscosity (mPa.s) 0.547
Water viscosity (mPa.s) 0.4554
Five wettability distributions are studied: f=0, f=0.25, f=0.5, f=0.75 and f=1. For the water-wet
case, as expected, water remains in the smallest portions of the pore space, giving very low
water relative permeability and significant trapping of oil in the larger pores at the end of
waterflooding, mainly caused by snap-off. In the case of poorly connected carbonates (ME1,
Portland and Indiana limestones), up to 75% of the pore space can be trapped. However, for
the better connected networks, namely ME2, Guiting and Mount Gambier, the water relative
permeability is higher and there is less trapping (there are more pathways for the oil to
escape), although the residual saturation is around 40% or higher in all cases.
1 1 1
(a) (b) (c)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1 1
(d) (e) (f)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Sw Sw Sw
Waterflood relative permeability for the strongly water-wet case (f=0). Curves
are presented in order of increasing connectivity. (a) Middle Eastern sample 1.
(b) Portland limestone. (c) Indiana limestone. (d) Middle Eastern sample 2. (d)
Guiting carbonate. (f) Mount Gambier limestone.
For a mixed-wet case with f=0.25, the small fraction of oil-wet pores tends to increase the
amount of oil trapping, particularly in the less connected networks where now there is little
or no range of saturation when two phases flow simultaneously, except very slow flow in
wetting layers. The water phase connectivity is reduced and the water relative permeability is
in general lower than the strongly water-wet case. The water-wet regions fill first in a capillary-
controlled displacement at the pore scale: these are the small pores and poorly connected;
however, they surround most of the oil-wet pores that are then trapped. These pores cannot
then be displaced during forced water injection, which explains the increase in residual oil
saturation. Here again, for the highly connected networks, the water relative permeability is
higher since the water has more possible pathways through the system and there is both
spontaneous and forced displacement by water.
1 1 1
(a) (b) (c)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1 1
(d) (e) (f)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sw sw sw
Waterflood relative permeability for the mixed-wet case (f=0.25). Curves are presented
in order of increasing connectivity. (a) Middle Eastern sample 1. (b) Portland limestone.
(c) Indiana limestone. (d) Middle Eastern sample 2. (d) Guiting carbonate. (f) Mount
Gambier limestone.
When the fractional wettability is 0.5, an equal mix of water-wet and oil-wet pores, at low
water saturations, a similar behaviour is observed regardless of the connectivity of the pore
space. At the beginning of the water flooding, the water is still poorly connected and flows
only through the smallest water-filled pores and thin wetting layers of the pore space;
therefore, the water relative permeability is low. However, in an equal mix of water-wet and
oil-wet fractions of the pore space, depending on the connectivity, an important increase in
the water relative permeability is noticeable. After spontaneous imbibition, a significant
forced displacement of oil occurs as the oil-wet pores and throats connect through the
network. The residual oil saturation is generally lower since oil remains connected in the oil-
wet region in layers. This effect is noticeable in the shape of the oil relative permeability, for
the well-connected samples, that show a long region where the oil relative permeability is
very low, but there is still displacement – this behaviour is controlled by slow flow in oil layers.
The poorly connected samples still show a water-wet controlled behaviour, where there is a
sharp decrease in the oil relative permeability and significant trapping. Here there is little
connectivity of the oil-wet regions and as a consequence layer drainage is unable to achieve
low residual saturations. In addition, the maximum water relative permeability varies from
very low to very high values dependent on the degree of trapping and the connectivity of the
water phase. Where the residual saturation is low, water can fill most of the pore space – and
the larger pores in the oil-wet regions – and has a high end-point value. A wide range of
behaviour is seen in this case dependent on the pore structure of the medium.
When the oil-wet fraction is higher, f=0.75, the residual saturation is now very low as the oil
remains connected in layers throughout the displacement. The water relative permeability
can rise to high values in all cases as the water fills the centres of the larger regions of the pore
space. This is a sign of a more typical oil-wet behaviour with displacement over a wide
saturation range and low relative permeabilities of both oil and water at low saturations of
their respective phases, controlled by wetting layer flow. This behaviour is generically similar
to network modelling calculations for sandstones (Valvatne and Blunt, 2004; Zhao et al.,
2010). The jumps in some of the curves reflect the relatively small size of the networks studied:
improvements in imaging should soon allow larger networks to be constructed.
1 1 1
(a) (b) (c)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1 1
(d) (e) (f)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sw sw sw
Waterflood relative permeability for the mixed-wet case (f=0.5). Curves are
presented in order of increasing connectivity. (a) Middle Eastern sample 1. (b)
Portland limestone. (c) Indiana limestone. (d) Middle Eastern sample 2. (d)
Guiting carbonate. (f) Mount Gambier limestone.
1 1 1
(a) (b) (c)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1 1
(d) (e) (f)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sw sw sw
For the fully oil-wet case (f=1) the behaviour is generally quite similar to that observed for the
mixed-wet case (f=0.75): very low residual oil saturation, a prolonged layer drainage regime
(low oil relative permeability at low oil saturation) and high end-point water relative
permeability.
1 1 1
(a) (b) (c)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1 1
(d) (e) (f)
0.8 0.8 0.8
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sw sw sw
Waterflood relative permeability for the strongly oil-wet case (f=1). Curves are
presented in order of increasing connectivity. (a) Middle Eastern sample 1. (b)
Portland limestone. (c) Indiana limestone. (d) Middle Eastern sample 2. (d)
Guiting carbonate. (f) Mount Gambier limestone.
To summarize the previous description, we analyze the impact of wettability and average
coordination number on the relative permeability behaviour. The evolution of residual oil
saturation with the fractional wettability shows that the residual oil saturation reaches a
maximum for the fractionally-wet case with f=0.25, and then decreases sharply to very low
saturations as the medium becomes more oil-wet. Waterflooding gives a high local
displacement efficiency for the cases f=0.75 and f=1, where the behaviour is controlled by oil
layers.
1
Guiting
0.9 Indiana
Portland
0.8 Mt Gambier
ME 1
0.7
ME 2
0.6
S
or 0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fractional Wettability
1
f=0
0.9 f=0.25
f=0.5
0.8 f=0.75
f=1
0.7
0.6
S
or 0.5
0.4
0.3
0.2
0.1
0
2 3 4 5 6 7 8
Average Coordination Number
The impact of connectivity on the residual oil saturation is shown above. The residual oil
saturation tends to decrease with increasing connectivity, regardless of wettability.
permeability which suggests that the cross-over point is at more than 50% water saturation
for water-wet cases and less than 50% water saturation for mixed-wet or oil-wet samples
(Craig, 1971). We only see this trend in the near oil-wet region; this rule does not apply in
general because of the low estimated water relative permeability.
We will now compare our computations to measurements found in the literature on reservoir
carbonate samples. The approach is not necessarily genuinely predictive as scans of the
reservoir samples and an independent measurement of wettability are not available: we
simply make an assessment if the estimated connectivity and wettability are plausible for the
experimental sample studied. Also, the objective of this comparison is not to have a perfect
match between the laboratory measurements and the results of the network modelling by
fine-tuning the oil-wet fraction or the contact angles; rather, the goal is to determine if our
calculated behaviour is supported by the available experimental evidence and discuss the
impact of wettability and pore structure on field-scale recovery.
Wettability NMR
Wettability Geology Lithology
measurement Description
Kharaib Dual pore Multi-modal,
Al-Sayari, 2009 Mixed-wet N/A
formation system microporosity
Mixed-wet
Meissner et al., 2009 Arab-D Lime
preference USBM
reservoir grainstone
to oil
Mixed-wet
Arab-D Lime Complex
Meissner et al., 2009 preference USBM
reservoir mudstone multi-modal
to oil
pore
Mixed-wet
Arab-D Lime structures,
Meissner et al., 2009 preference USBM
reservoir grainstone Microporosity
to oil
Mixed-wet
Arab-D Lime
Meissner et al., 2009 preference USBM
reservoir grainstone
to oil
Arab-D
Neutral to
reservoir
Okasha et al., 2007 slightly Amott N/A N/A
Haradh
water-wet
area
Generally Static Arab-D
oil-wet to imbibition reservoir
Okasha et al., 2007 N/A N/A
intermediate- Amott Utmaniyah
wet USBM area
A summary of the petro-physical and geological descriptions of the reservoir
samples found in the literature. USBM stands for US Bureau of Mines and is a
somewhat cumbersome method to measure wettability: it measures the ratio of
the area under the capillary pressure curves for spontaneous and forced
displacement.
1 1
(a) (b)
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
Kr
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Sw Sw
A similar relative permeability to that measured can be observed for the case of f=0.25 for the
well-connected Guiting and Mount Gambier networks. The relatively low residual oil
saturation and the shape of the oil relative permeability curve indicate a mixed-wet behaviour.
For Guiting, the discrepancies in the water relative permeability can be explained by the un-
resolved microporosity.
Case 2. Meissner et al. (2009) performed detailed measurements on several samples from the
Arab-D reservoir of the Dukhan field, onshore Qatar. They reported the results of several of
steady-state relative permeability tests for oil/brine and gas/oil systems. Results were
reported for both native and the restored state cores. The results were reported in terms of
normalized saturations and relative permeabilities:
−
=
1− −
(16.8)
where Swi, the initial water saturation, is determined after primary drainage and Sor is the
residual oil saturation determined by extrapolation of the oil relative permeability as it
asymptotically approaches zero.
In this case, to introduce an initial water saturation, we set the maximum primary drainage
capillary to be equal to 690 kPa (approximately 100 psi). This value is chosen based on the
different capillary pressure measurements that showed a sharp increase in the pressure for
an average pressure of around 100 psi.
The figure overleaf shows a comparison between the four measurements reported of
water/oil relative permeability on the native state subsurface cores with the relative
permeability generated for the strongly oil-wet case f=1 for ME1. The suggestion here is that
the reservoir is strongly oil-wet with a structure similar to that observed in the subsurface
sample from which we extracted a network.
1 1
Kr ( frac. K o,swi)
0.8 0.8
0.6 0.6
0.4 0.4
(a) (b)
0.2 0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.5 1
1 1
Kr ( frac. K o,swi)
0.8 0.8
0.6 0.6
0.4 0.4
0 0
0 0.2 0.4 0.6 0.8 1 0 0.5 1
Swn Swn
A comparison between the w ater flood relative permeability for the Middle
Eastern sample 1, for a strongly oil-wet case f=1 w ith measurements on native
state subsurface reservoir cores (oil relative permeability, circles, and water
relative permeability, crosses) obtained from Meissner et al. (2009).
The figure overleaf shows a good agreement between the measurements and the relative
permeability generated by network modelling for the mixed-wet Mount Gambier network
(f=0.25) for one of the three samples. Note that we suggest that in this field the wettability
and pore structure are different from the subsurface Middle Eastern sample in the previous
section. The figure below shows good agreement for the second measured sample with low
connectivity carbonates i.e. Portland and ME1 for a strongly oil-wet case. The difference of
the wettabilities is evidence of local variations of wettability within the reservoir.
Good agreement was not obtained for the third sample which had high connate water
saturation.
0.9
0.8
0.7
0.6
kr 0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
sw
Mount Gambier water flood relative permeability for the mixed-wet case with an
oil-wet fraction of f=0.25 (solid) compared to measurements on a reservoir
sample obtained from Okasha et al. (2007) (oil relative permeability, circles, and
water relative permeability, crosses).
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
kr
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
sw sw
Middle Eastern sample 1 (right) and Portland limestone (left) waterflood relative
permeability for the strongly oil-wet case w ith an oil-wet fraction of f=1 (solid)
compared to measurements on reservoir samples obtained from Okasha et al.
(2007) (oil relative permeability, circles, and water relative permeability,
crosses).
In waterflooding, for oil and water of similar viscosity, the saturation at which the relative
permeabilities cross – as discussed below – gives a useful and simple indicator of the recovery.
For water saturations beyond the crossover point, more water will be produced than oil –
beyond this point, oil production becomes increasingly uneconomic. Hence, a rough guide to
recovery can be derived from the change in saturation from its initial value to when the
relative permeabilities cross.20 Later, we show how to perform this analysis rigorously and
predict – for given relative permeabilities and fluid viscosities – the amount of oil recovered
as a function of the amount of water injected.
Our simulations indicate that the optimal water flood efficiency is observed for a mixed-wet
system with a large fraction of oil-wet pores, around 0.75. The highest waterflood efficiency
20 Please note, that this change in saturation is not a recovery factor. For this, we need to
compute the volume of oil produced, convert it to surface conditions and then divide by the
total volume of oil initially in the reservoir (also measured at surface conditions). Last this is
an approximate physically-motivated assessment of recovery and no excuse for not doing a
proper analysis, as described later.
is implied for the less well-connected samples, since in these cases the waterflood relative
permeability is very low and this holds back the movement of water, allowing oil to be
displaced. For better connected samples, there is less sensitivity to wettability and overall a
lower crossover saturation, indicating less favourable recoveries.
This is a somewhat surprising conclusion and implies that waterflooding in mixed to oil-wet
carbonates of poor pore-space connectivity may be an effective process. This behaviour
stands in contrast to sandstones, where network modelling studies indicate that more
neutrally-wet conditions provide optimal recovery (Øren et al., 1998; Valvatne and Blunt,
2004). Moreover, experimental measurements presented by Jadhunandan and Morrow
(1995) have shown that oil recovery by waterflooding in sandstones reach a maximum at close
to neutral wettability.
As mentioned at the beginning of this section there are two distinct recovery processes in
carbonates, depending on whether or not fractures dominate the flow. If they do not, then
viscous forces are significant for displacement through the porous matrix and local recovery
is determined by the relative permeabilities. It is possible to perform a Buckley-Leverett
analysis to compute, analytically, recovery for a homogeneous one-dimensional displacement
from the relative permeabilities: the method to do this is presented later in these notes
(section 19). However, the as mentioned previously, the likely local waterflood displacement
efficiency can be estimated rapidly from direct inspection of the relative permeability curves.
Imagine that the reservoir-condition oil and water viscosities are the same. Then, if the
saturation near the production well is where the relative permeabilities cross, the subsurface
ratio of oil to water production will be 1:1. Wells are abandoned when the cost of recycling
and processing the produced water exceeds the economic benefit of the oil produced: this is
normally when the oil:water ratio is between 1:2 and 1:10. On the other hand, the oil viscosity
is typically greater than that of water, and the flow rate is determined by the ratio of relative
permeability to viscosity. Hence, in most cases, production ceases close to where the relative
permeabilities cross – between the producer, where water has displaced oil, the saturations
will be higher, but this very simple trick allows a quick comparative study of recovery trends.
Hence, waterflooding is quite favourable in the less well connected carbonate samples. Most
of the moveable pore volume is displaced, and the residual saturation is low. The reason for
this is that the poorly connected water phase holds back water advance, allowing the efficient
displacement of oil. For better connected samples, the water is better connected and rapidly
finds a pathway of large pores through the system. This allows water to bypass oil at the pore
scale, leading to less favourable waterflood recovery.
Now consider a reservoir where flow is dominated by fractures. In this case the fractures
effectively short-circuit the flow field and it is not possible to impose a substantial viscous
pressure drop across the matrix. Recovery is mediated by capillary and gravitational forces.
Imagine that water quickly invades the fractures surrounding a region of matrix (a so-called
matrix block, although it does not have to be exactly, or even remotely, cuboidal in shape).
Then recovery will occur by spontaneous imbibition – that is recovery will occur until the
capillary pressure is zero. Shown below are the capillary pressures for three carbonate
samples. These are not the same samples as already discussed, with the exception of Mount
Gambier, with a well-connected pore space. Ketton and Estaillades are both more poorly
connected and are – for the sake of this discussion – similar in behaviour to the other low
coordination number samples, Portland, Indiana and ME1.
Primary drainage (solid line) and waterflood (dotted line) capillary pressures
predicted using pore-scale netw ork modelling. Here the oil-w et fraction f is 0.75.
If recovery occurs by imbibition – in a fractured reservoir – the final recovery is
controlled by the saturation when the capillary pressure is zero, not the residual
saturation.
In our examples in the figures above this means that only around 25% of the moveable pore
volume is recovered. Furthermore, the rate of recovery is limited by the rate at which water
can advance into the pore space – that is the relative permeability in the low water saturation
range where the capillary pressure is positive. In this case now the more favourable system
is the Mount Gambier – the water relative permeability is higher, indicating a more rapid
displacement, while the degree of spontaneous imbibition is larger, since the well-connected
pore space allows all the water-wet regions of the rock to be accessed easily; in contrast the
poorly connected Ketton and Estaillades has a lower water relative permeability and not all
the water-wet regions of the pore space are interconnected, leading to less displacement at
a positive capillary pressure.
Gravitational forces can also play an important role in the displacement. If water floods a
vertical fracture then oil, being less dense, is preferentially produced from the top of the
matrix, then the weight of water in the fracture acts as a driving force. If we assume that the
capillary pressure in the fractures is very small and is equal to zero at the top of a matrix block,
then the capillary pressure at the base is gh, where is the density difference between
water and oil and h is the effective height of the matrix block. The capillary pressure is
negative: the water has a higher pressure than oil. This allows forced displacement to a lower
oil saturation. Taking typical values: g = 9.81 ms-2; = 300 kg.m-3 and, say, h = 2 m, then the
negative capillary pressure that can be reached is around -6 kPa. Reading off the graph above,
we can see that this driving force displaces a further 15% of the oil for the lowest permeability
sample, Estiallades. Even if we consider lower permeability rocks (the capillary pressure
approximately increases as 1/K1/2, where K is the permeability – this is a consequence of
Leverett J-function scaling, section 10), there is likely to be significant displacement with this
driving force and demonstrates how both capillary and gravitational forces mediate recovery
in field settings.
Gravity also determines the initial water saturation before waterflooding. As is apparent from
the capillary pressures, for the lowest permeability sample, Estaillades (which is still high
permeability compared to most reservoir rocks) an effective matrix block height of around 10
m would be required to displace all the oil to close to residual saturation. There is a corollary
to this: it also indicates that the initial saturation determined by capillary-gravity equilibrium
(based, typically on the primary drainage capillary pressure) has a transition zone – with
varying saturation above the irreducible value – of height around 10 -100m for rocks with
permeabilities between 1mD and 100mD (using the 1/K1/2 scaling mentioned above). The
initial water saturation affects both the wettability (at high saturation less of the rock is
contacted directly by oil and, as the imposed capillary pressure is lower, the wettability
alteration is likely to be less strong) and the starting point for waterflooding. There is often a
wettability trend from water-wet near the oil-water contact, through mixed-wet in most of
the reservoir with more oil-wet conditions at the crest, as discussed previously. Usually,
coreflood measurements are made from samples near the top of the reservoir: this could
suggest oil-wet conditions and unfavourable waterflood recovery, when the reality is a much
more efficient displacement in most of the reservoir column. Pore-scale modelling, allowing
the prediction of relative permeabilities as a consistent function of initial water saturation,
has enormous potential to improve the characterization of such reservoirs.
This rather simple analysis already leads to some interesting and surprising conclusions. For
the same wettability, in a reservoir where flow is not fracture dominated, local waterflood
recovery is higher in the lower-permeability less well-connected sample, since the low water
relative permeability holds back the water advance. On the other hand, if the reservoir is
extensively fractured, the better-connected sample gives faster and better recovery, since
there is a greater degree of spontaneous imbibition allowed. This is a clear indication that
both the nature of the reservoir – fractured or unfractured – and the multiphase flow
properties are both crucial for any reasonable assessment of recovery.
I call this conundrum over recovery – mixed-wet systems are good for displacement, but bad
for imbibition – the ‘trillion barrel question’ since it will determine the recovery of most of our
remaining conventional reserves of oil, mainly in the Middle East. While I do not have a simple
answer – and it is unlikely that there is a simple answer – it does underscore the importance
of this topic and how important it is to have good measurements (and predictions) of relative
permeability.
Oil, water and gas may all flow together in reservoirs. Examples include gas injection,
including carbon dioxide injection, solution gas drive (when the reservoir pressure is dropped
below the bubble point), gas cap expansion and steam injection. In environmental settings,
when a non-aqueous phase pollutant migrates downwards towards the water table in a moist
soil, there are three mobile fluid phases – the pollutant, water and air.
First we will consider oil, water and gas at the pore scale, and how they are arranged. We use
this insight to discuss wettability and relative permeability, as well as oil recovery.
Gas go
gw
Oil
Water
ow
The arrangement of a small droplet of oil floating on water in the presence of a gas.
= − −
(17.1)
If Cs > 0, the oil spreads on water. Light alkanes and many alkane mixtures – and indeed most
crude oils – spread: the arrangement shown in the figure above is not stable and it will be
energetically favourable for the oil will cover the interface between gas and water.
If Cs < 0, the oil does not spread on water. Dense (denser than water) non-aqueous phase
liquids (such as chlorinated solvents) and long chain alkanes (such as decane, dodecane etc)
do not spread on water.
In thermodynamic equilibrium, there are then three things that can happen to the drop of oil.
Gas
Oil
Water
The equilibrium arrangement of a small droplet of oil floating on water in the presence
of a gas. Here the spreading coefficient is negative and no oil film forms between the
gas and water.
Gas
Oil
Water
The equilibrium arrangement of a small droplet of oil floating on water in the presence
of a gas. Here the initial spreading coefficient is positive; an oil film forms, generating
a new effective gas/water interface with a negative equilibrium spreading coefficient.
3. Cs > 0 and Cse = 0. The oil film swells without limit as more oil is added to the system.
When the oil film is sufficiently thick, the effective interfacial tension between the gas
and water is the sum of the interfacial tensions between oil and water, and gas and
oil. This means that the equilibrium spreading coefficient is zero.
Gas
Water
The third and final possibility for the drop of oil is that it spreads without limit forming a
thick film. Effectively the interfacial tension across the interface containing the film is
simply the sum of the oil/water and gas/oil interfacial tensions, giving an equilibrium
spreading coefficient of zero.
then oil would continue to spread until its value was zero – case 3 above.
Now we will discuss what this implies about the arrangement and flow of three phases in the
pore space. The typical arrangement of the phases is shown in the figure below.
Water
Oil
Gas
The arrangement of oil, water and gas in a water-wet pore of triangular cross-section.
Note that the oil resides as a layer sandwiched between the water in the corners and gas
occupying the centre of the pore.
The oil can form layers in a (water-wet) pore space, sandwiched between water in the corners
and gas in the centre. This formation of the oil layer is favoured by having a low (or zero)
equilibrium spreading coefficient, as this controls the contact angle between the gas and oil.
However, for an oil layer to form – that is to be able to draw a layer in an angular pore space:21
+ /2
(17.2)
Spreading oils (with an equilibrium spreading coefficient of zero) have go = 0 since, at the
microscopic level, there is no angle of contact between the gas and oil. go increases as the
spreading coefficient becomes more negative, as we will show below.
21 This is a necessary, but not sufficient condition for the formation of a layer. Strictly, we need
to consider an energy balance for the displacement. Conceptually this is straightforward, as
we consider the change in interfacial energy associated with different possible fluid
configurations, but can be complex when dealing with three-phase flow.
Consider the Young equation on a flat surface, with different combinations of fluids, shown in
the diagrams overleaf. Re-arranging the equations leads to an important equality in three-
phase flow, known as the Bartell-Osterhof (1927) equation:
= +
(17.3)
Here we assume that the contact angles and interfacial tensions are measured in
thermodynamic equilibrium. This relationship provides a constraint between the contact
angles and the interfacial tensions: there are only two independent contact angles.
Conventionally we consider that the wettability controls ow, while spreading controls go.
If the system is strongly water-wet with gw = ow =0, then the gas/oil contact angle ow is
simply:
=1+
(17.4)
using our definition of spreading coefficient, which has a negative (or zero) value.
Now consider that we have residual oil surrounded by water. Now consider that gas enters
the system. Examples of this include lowering the water table in a soil with non-aqueous
phase pollutants present, gas injection in oil reservoirs, solution gas drive (the primary
production mechanism that occurs when the pressure in an oil field drops below the bubble
point) and gravity drainage through gas cap expansion. In a water-wet system, this oil will
occupy the centres of the larger pore spaces.
When the gas phase is introduced, oil spreads in the porous medium between water (coating
the solid surfaces) and gas (which as the most non-wetting phase preferentially fills the
centres of the largest pores). Oil layers for that occupy the crevices and corners of the pore
space between water and gas – it is now connected wherever there is gas, and the oil can
flow. We can drain to essentially zero saturation. Remaining oil saturations as low as 0.1%
have been observed after gravity drainage in sand packs (Sahni et al., 1998).
A full discussion of three-phase flow rapidly becomes rather complicated, beyond the key
concept of oil layers: as an example the figure on the next two pages shows (just some of) the
configurations of two and three phases in a pore space, dependent on saturation path and
wettability. All of these fluid configurations simply use the concepts of wetting, spreading,
contact angle and the Young-Laplace equation. Rather than go through all the details, we will
present, briefly a discussion on wettability, pore-scale configurations (specifically layers) and
recovery later in this section. However, we can also see oil layers in mixed-wet and oil-wet
media; gas is always non-wetting to oil, and so can reside in the centre of a pore, with an oil
layer (as in two-phase flow) present in the corners. However, as we show later, gas is not
necessarily non-wetting to water in an oil-wet system.
(This and the previous page) Some of the possible configurations of three
phases – oil, w ater and gas – in a corner of the pore space, taken from Piri and
Blunt (2005).
We expect to find very low values of the residual oil saturation in the presence of water and
gas, because of oil layers, as discussed above, but the relative permeability will be low as well.
Three-phase relative permeability is extremely difficult to measure. There are also a huge
range of different saturation paths that may be taken during a displacement that all may have
different relative permeabilities.
Normally three phase relative permeabilities are predicted using empirical models with a
dubious physical basis. A full discussion of the various models is outside the scope of these
notes. In a water-wet system, the water tends to reside in the small pores, the oil in the
intermediate pores and the gas in the large pores, consistent with our pore-scale picture. This
means that the exact pore sizes seen by the oil depends on the amount of water and gas
present in the porous medium, leading to relative permeabilities that depend on two
independent saturations. If the system is mixed-wet this picture is further complicated. In
general, in three-phase flow – in the presence of gas – the oil relative permeability is lower,
making recovery low, but the residual oil can also be very low, leading to high ultimate
recoveries, because of the drainage of oil layers. The gas relative permeability – in a water-
wet system – is high, since gas resides in the larger pores, leading to early breakthrough and
poor overall recovery when gas is injected. On the field scale, the main design criterion is how
to keep the injected gas in the reservoir, allowing the oil to flow to low saturation. It is less
easy to make general statements about recovery, as the system is now much more complex,
and we have to rely on field-scale simulation models to assess recovery efficiency.
Trapped oil and gas imaged in a w ater-wet sandstone (Iglauer et al., 2012).
Initially the core is full of water. Then oil is injected – this is primary drainage.
Then two displacement sequences are considered. The first is gas injection,
followed by water injection (gw), while the second is w aterflooding, follow ed by
gas injection followed by a further waterflood (wgw). The gw sequence leads to
considerably more trapping of gas in the pore space.
The amount of trapping – of both gas and oil – is dependent on the displacement sequence,
with less trapped if we waterflood the reservoir before gas injection (this is called tertiary
injection, as opposed to gas injection straight away, which is secondary gas injection). The
morphology of the trapped clusters is also different, with smaller clusters of gas seen for the
wgw sequence. This is an active topic of research and we do not have, as yet, a full
understanding of recovery and displacement processes in three-phase flow.
One important observation is that – in water-wet systems – more gas can be trapped in a
three-phase displacement than if displaced only by water. This could be used to design gas
injection to retain the gas in the reservoir, while mobile oil is produced. Experimental
evidence for this is shown below for experiments on sand packs: more gas can be trapped in
a three-phase displacement involving oil and water than when gas is displaced by water alone.
In general, there is more trapping of oil and gas combined than in two-phase flow, more
trapping of gas, but less trapping of oil alone.
20
18
16
14
12
Sgr(%) 10
8
6 Drainage time 30min
Drainage time 17hrs
4
Drainage time 2hrs
2 Drainage time 24 hrs
0
0 20 40 60 80 100
Sgi(%)
If we consider all the various possible configurations of phases in the pore space, and have a
good network representation of the rock, it is possible to make predictions of three-phase
relative pemeabilities. The dataset is the classic Berea measurements of Oak (1990) compared
to network model predictions of Piri and Blunt (2005). Overall, bearing in mind the complexity
of the problem, the predictions shown are a good test of the ability of pore-network modelling
to predict the behaviour of complex systems.
Measured and predicted gas relative permeabilities. The different points refer
to different displacement sequences in the experiments. From Piri and Blunt
(2005).
We will now make some statements concerning wettability, layers and recovery. The figure
below shows the results of gravity drainage experiments (gas enters a long sand column, while
oil and water drain out of the bottom of the column under gravity the same process occurs in
an oil reservoir, if gas is introduced to the crest of the field, or a natural gas cap expands). The
relative permeability is shown for three cases: octane as the oil in a water-wet system; decane
as the oil in a water-wet system; and the water relative permeability for an oil-wet case.
Measured relative permeabilities for gas gravity drainage in a sand pack. From
left to right: the oil relative permeability in a water-wet medium, when octane is
the oil phase; the same experiment but with decane as the oil; the water relative
permeability in an oil-wet system. From DiCarlo et al. (2000).
The behaviour is different in each case and we can understand this and discuss the
implications for recovery using our discussion of spreading coefficient and the Bartell-
Osterhof relation, Eq. (17.3).
Octane spreads – or almost spreads – on water, with an effective gas/oil contact angle in a
water-wet system close to zero. Hence, oil layers readily form. If gas is injected, oil layers
form and allow drainage down to very low saturation – below 1% in the sand pack studied. At
low oil saturation, the flow is dominated by this layer drainage. The oil saturation is simply
proportional to the area of oil open to flow. The conductance scales as the square of the area
– this is important – the square of the area, rather than proportional to it. This is a direct
consequence of the Navier-Stokes equation: consider Poiseuille flow where the flow rate is
proportional to the fourth power of radius (second power of area). Physically, this is because
there is no flow at a solid boundary. If we increase the area to flow, then the flow speed in
the centre of the channel can increase, as it is further away from the walls. This, combined
with the fact that the area is greater, is what leads to the quadratic dependence on area. This
result contrasts with electrical conductance, which scales linearly with area.
The oil relative permeability is just the fractional flow conductance of the oil and so this
discussion leads to the prediction:
~
(17.5)
Decane has a higher interfacial tension with water than octane and does not spread on water.
Its contact angle in the presence of gas is non-zero and this non-spreading oil does not form
oil layers in the pore space. Hence there is no oil layer drainage regime, the oil relative
permeability drops rapidly at low saturation and we see significant trapping. This is observed
in the middle figure above.
It is considered likely that in reservoir settings, most oils are spreading, and so high recoveries
are potentially possible. However, in environmental applications many non-aqueous phase
liquids, particularly chlorinated solvents, do not spread on water and therefore can remain
trapped even in the presence of gas (air).
If instead the porous medium is oil-wet, then perhaps we can simply swap phases. That is,
the water relative permeability in an oil-wet system is the same as the oil relative permeability
in a water-wet system. The right-hand figure for the gravity drainage experiments shows that
this is not the case if the oil is spreading. The water relative permeability drops sharply and
has an irreducible saturation. Why is this, and what has it got to do with ducks?
The figure below explains this. If we re-arrange the Bartell-Osterhof equation to find the
gas/water contact angle for a spreading oil (go =0) in a strongly oil-wet system (ow = 180o),
we find:
−
=
(17.6)
The interfacial tension between oil and water is always larger than between gas and oil. Hence
the gas/water contact angle has a negative cosine and is greater than 90o. Water cannot
spread on oil in the presence of gas. Indeed, in a strongly oil-wet system water is the most
non-wetting phase and can be trapped by gas and water.
Not
allowed
Diagram showing oil layers in the pore space of a water-w et medium (left). In
contrast, water layers cannot form in an oil-wet system, as the gas/water contact
angle w ill be greater than 90 o .
This is why ducks don’t get wet. Their feathers are covered in oil and form an oil-wet porous
medium. This makes water the most non-wetting phase: you have to force water into the
feathers – they prefer to be surrounded by air keeping the duck dry and insulated. This is also
why water beads and runs off an oily surface. It also why sea-bird often die if crude oil (from
a spill) is washed off their feathers using soap: now their feathers are water-wet, they imbibe
water and the birds soon die of hypothermia.
While this observation is the topic of a children’s book (see the next page), the concept still
struggles to be accepted by petroleum engineers, where it is still widely assumed that gas
‘must’ be the most non-wetting phase.
The consequences of wettability for recovery in three-phase flow are still not fully understood.
We do expect good ultimate recoveries if oil layers form – as they do in a spreading system,
regardless of wettability (oil is always more wetting that gas). We can also trap and suppress
the movement of gas in mixed-wet media, which is favourable for recovery.
You have got near the end of a demanding and difficult set of notes and now
you know as much as ….,. a young child who has read this book!
The approach so far in the notes has been physically motivated with a minimum of equations.
However, to provide some rigour and background to our comments on recovery, we will now
derive flow equations for multiple phases in a porous medium.
We will consider a conservation equation for a case where we have multiple phases. This is
an extension of the derivations for single-phase flow in section 14. The approach will be very
slightly different, but is straightforward. Here we consider the transport of saturation, rather
than concentration. Consider conservation of mass of one phase (water) in the diagram
below.
Flow of water
x x+x
x
The mass of water that enters the box – the shaded region shown in the figure above – in a
time t = At w qw(x), where qw is given by the multiphase Darcy law, Eq. (16.1). A is the
cross-sectional area to flow. Similarly, the mass that leaves is given by At w qw(x+x). The
mass of water in the box is given by AxSw. The mass in minus the mass out is the change in
mass:
∆ ( )− ( +∆ ) = ∆ ( +Δ )− ( )
(18.1)
( +Δ )− ( ) ( +∆ )− ( )
+ =0
∆ ∆
(18.2)
Where we have assumed that the density and porosity are constant (assuming, as in section
14, that the flow is incompressible). Then we take the limit of small x and t to obtain a
differential equation:
+ =0
(18.3)
This simple form of the conservation equation can be rearranged by substituting in Darcy’s
law, Eq. (16.1). This takes some algebra and ends with an equation that can be solved
analytically for saturation. We start with Eq. (16.1) for one-dimensional flow of the water
phase:
=− −
(18.4)
=− −
(18.5)
and Pc = Po – Pw is the capillary pressure. kro, krw and Pc are known as functions as Sw.
+ =0
(18.6)
Add the two conservation equations (18.3) and (18.6):
( + ) ( + )
+ =0
(18.7)
We define qt = qw+qo as the total velocity. Then Eq. (18.7) is (the saturation term is zero
as for two-phase flow, the sum of the oil and water saturation is one, a constant):
=0
(18.8)
The total velocity is constant in space (it can vary over time; qt(t)) for one-dimensional flow.
From the multiphase Darcy equations for oil and water, Eqs. (18.4) and (18.5) and writing the
expression in terms of the water pressure only:
= + =− − − + −
(18.9)
Then defining mobilities by w=krw/w and o=kro/o, with the total mobility given by t=wo,
Eq. (18.9) becomes:
=− + ( + )−
(18.10)
= − − + −
(18.11)
= + ( − ) +
(18.12)
In words the water Darcy velocity = pressure gradient + capillary pressure + gravity.
We can write the conservation equation (18.3) in terms of the water fractional flow – the
fraction qw/qt:
+ =0
(18.13)
= 1+ +( − )
(18.14)
The fractional flow has three terms representing the three physical forces that impact the fluid
movement: advection (governed by the pressure gradient); capillary pressure and gravity
(buoyancy).
Many authors define mobility as w=Kkrw/w. with an extra factor of K. Also we use Q (volume
per unit time) and q (Darcy velocity) rather than q and v respectively, as in some books. Often
you see conservation equations with an explicit area A. Our equations are per unit area: Q =
qA.
The conservation equations we have developed are for an oil/water displacement, where
water displaces oil, as occurs in a hydrocarbon reservoir.
We will now address some special cases, where simplifications to the equations can be made.
While we cannot solve the full equation directly, we can explore solutions in a variety of
different limits. One case will be considered in this section.
If we have gas/water flow, then the mobility of the gas phase can be considered to be much
larger than for the water. The Richard’s equation describes transport of water in this case.
In Eqs. (18.13) and (18.14) o is replaced by g (for gas) and g >> w and so t = g. Then the
fractional flow can be written as:
= +( − )
(18.15)
where the first – advection – term in Eq. (18.14) is now considered to be negligible. Then the
conservation Eq. (18.13) is:
+ +( − ) =0
(18.16)
More usually this equation is written in terms of the pressure head p = Pw/wg + z. Write =
Pw/wg. If the gas (air) density is considered to be negligible and the air pressure constant at
atmospheric (Pc = Patm – Pw) then Eq. (18.16) becomes for vertical flow (gx = g):
= −1
(18.17)
using the standard definition for hydraulic conductivity, KH (see section 12). Often rather than
seeing the relative permeability and capillary pressure written as a function of pressure, the
saturation and relative permeability are written as a function of scaled capillary pressure ( ).
Eq. (18.17) is the standard transport equation in hydrology to describe the movement of water
under gravity and capillary pressure.
We will now begin to construct an analytical solution for multiphase flow in one dimension.
The derivation can be somewhat cumbersome and is helped if we first develop the concept of
the fractional flow: the fraction of the total flow of oil and water that is taken by water alone.
g
gx=gsin
Consider a reservoir with constant dip and linear flow, as shown in the figure above. Using
the conservation equation (18.13), we can write the fractional flow Eq. (18.4) as:
= 1+ + sin
(19.1)
where is the angle to the horizontal and the density difference is written as .
How important is the capillary pressure term at the field scale? While it dominates at the pore
scale, as discussed previously in these notes, it is small over 100s m to km: the effect of
capillary pressure is encapsulated in the relative permeabilities. This was discussed in section
15 in the context of capillary and Bond numbers.
We will estimate the relative contribution of viscous forces (advection), capillary pressure and
buoyancy to the fractional flow by considering the magnitude of each of the three terms in
the brackets of Eq. (19.1). By definition the viscous term is 1. For qt of the order of 10-5 ms-1
(around 1 m/day), K of 10-13 m2 (100 mD), a viscosity of 10-3 Pa.s and a relative permeability of
order 1, then the term in Eq. (19.1) is around 10-5 m.Pa-1. Then for a typical capillary
pressure of 104 Pa varying over a typical distance between wells (say 100 m), the capillary
pressure term in Eq. (19.1) is of order 10-3. The gravitational term, for, say, a density difference
of 300 kg.m-3, is around 0.03. This indicates that buoyancy forces are generally small (but not
negligible) in comparison to advection, while the effect of capillary pressure – at the field scale
– is tiny. Another way of seeing this is to consider the pressure drop between injection and
production wells – generally a few MPa – compared to capillary pressures of 0.01 to 0.1 MPa.
So, to construct an analytical solution applicable at large scales, we ignore the capillary
pressure (later, in section 20, we will solve for the opposite limit – where capillary pressure
dominates – which applies for imbibition in fractured media) and write Eq. (19.1) as:
= 1+ sin
(19.2)
Define a gravity number (this is different from the Bond number introduced in section 15):
(19.3)
1 1
= =
1+ 1+
(19.4)
1+ sin
=
1+
(19.5)
Let’s look at typical fractional flow curves, shown overleaf, for = 0 as a function of the
endpoint water/oil mobility ratio,
(19.6)
The curves have a characteristic S shape, with a point of inflection when we consider
horizontal flow (without gravity).
M=10
fw
M=1
M=0.1
0
0 Swc 1-Sor
Sw
Example fractional flow curves for a horizontal system with different end-point mobility
ratios, M, Eq. (19.6).
If we include gravity, we can study the behaviour schematically for M=1, where =
. Here the fractional flow curves can be greater than one, or less than zero. This
represents, physically, counter-current flow, where water moves downwards while oil moves
up: if the overall flow direction is downwards and the oil flows upwards, then the water flow
is greater than qt and the fractional flow is greater than one. Conversely, if the flow is uphill
and water is flowing downhill, then the water fractional flow is negative. Mathematically, fo
+ fw = 1. Thus when fw>1, fo<0 and vice versa: the phases are flowing in opposite directions.
Hence, if gravity is sufficiently strong, it can generate counter-current flow of oil and water.
To repeat: this means that the water moves downhill, while oil moves uphill.
n=3
fw
n=0
n=-3
0
0 Swc Sw 1-Sor
Example fractional flow curves including the effect of gravity. If gravitational effects are
sufficiently strong, we can have – for some range of saturation – counter-current flow
where the oil and water flow in opposing directions. In this case, the water fractional
flow is either greater than 1 or negative.
We have now derived the conservation equation and plotted different typical fractional flow
curves. For reference, we will continue with a specific example, shown in the figure below.
The water and oil relative permeabilities are written as follows (this is a common
( − )
=
(1 − − )
(19.7)
( − )
=
(1 − − )
(19.8)
In this specific case we take a maximum water relative permeability of 0.5, a maximum oil
relative permeability of 0.8, a=4 and b=1.5 with Swc=0.2 and Sor=0.3. For the fractional flow
the water and oil viscosities are 0.001 and 0.03 Pa.s respectively: the relative permeabilities
and fractional flow are shown below. From Eq. (19.6) the mobility ratio M=18.75 and n=0
(horizontal flow).
We will now show how to solve the conservation equation (18.13) with the fractional flow
given by Eq. (19.5).
+ =0
(19.9)
+ =0
(19.10)
22 These are sometimes called Corey curves and the power-laws Corey exponents. In reality,
the original paper was a little more specific and proposed a physical justification (which has
no foundation) for presenting equations for relative permeability and did not directly present
these simple forms, but in any event Corey (or Brooks and Corey) were the first authors to
suggest fitting relative permeability data to a power-law form.
We are interested in water injection from a well into a reservoir containing some initial
(usually irreducible) water saturation. We consider one-dimensional flow with an injection
well placed at x=0 and a producer at x=L. Then the initial condition is: t=0, Sw(x,0) = Swi, while
the boundary condition at the well is x=0, (well) Sw(0,t) = Sw0.
We control rates (fw) at wells, not saturation. Thus we find Sw0 that has given fw(0,t). Normally
we have Swi = Swc and inject 100% water: hence fw(0,t) = 1, and Swo = 1 – Sor.
Relative permeabilities
1
0.9
0.8
0.7
Realtive permeability
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Water saturation
Relative permeabilites used for our example test case. This represents a weakly water-
wet system.
0.9
0.8
0.7
0.6
Fractional flow
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Water saturation
The water fractional flow corresponding to the relative permeabilities shown previously.
(19.11)
We also define a dimensionless time. This requires some more thought – tD is the pore
volumes of water injected. This gives an indication of how much water has entered the
system, in comparison with the total capacity of the reservoir. It is defined as follows:
1
= = = =
(19.12)
where Q is the total flow rate and Vp = AL is the pore volume. If the flow rate is constant,
then the integral simply becomes Qt: however, this definition does allow varying flow rates to
be accommodated in the analysis. A useful relationship, which we will use later is to convert
dimensional speeds to a dimensionless quantity. If we have a speed v=x/t, then using
Eqs.(19.11) and (19.12), v=qt/ vD, where vD=xD/tD.
Then transformation of variables means that the conservation equation (19.10) becomes:
+ =0
(19.13)
We will now solve this equation by the method of characteristics (MOC). What this means is
that we will find the dimensionless velocity with which a given saturation moves. This means
that we find the solution as a function of dimensionless velocity. At first, this can be a
confusing concept. However, for a given time, the profile of saturation as a function of velocity
is the same shape as saturation as a function of distance: the profile elongates linearly with
time. It is also straightforward to convert between dimensionless and real variables, with
care.
= =−
(19.14)
1
= =
(19.15)
23 An entirely equivalent approach is to assume that the solution for water saturation is a
function of z=x-vt only and find v. This makes it even more explicit that the solutions are
profiles which move with some characteristic speed.
− =0
(19.16)
One solution is a so-called constant state – that is a saturation that does not change with
dimensionless wavespeed. The non-trivial solution is:
(19.17)
Let’s look at the example fractional flow shown previously: since vD=dfw/dSw we find the
derivative which resembles the curve shown overleaf.
Re-arranging the plot, to find water saturation as a function of speed (which – at a fixed time
– would represent water saturation as a function of distance) we arrive at something that
does not make sense – we have multiple solutions, as shown in the plot below.
5
Fractional flow derivative
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Water saturation
A schematic of the derivative of the fractional flow. Note that the derivative – indicating
dimensionless wavespeed – has a maximum at intermediate water saturation.
0.8
0.7
0.6
0.5
Water saturation
0.4
0.3
??? How do we make
sense of this?
0.2
0.1
0
0 1 2 3 4 5 6 7
Derivative of the fractional flow (dimensionless velocity)
19.2 SHOCKS
Shocks are discontinuities in saturation, which means that we can’t use a differential equation
to describe them. Shocks are encountered in other physical situations, such as a bomb blast,
the flash of light after a nuclear explosion, or traffic jams.
Consider the situation overleaf with a shock between two saturations – a left state and a right
state. We are only showing the region at the shock – the saturations can vary smoothly with
distance away from the shock.
SwL-SwR
Sw
vsh
dx=vshdt
Imagine a shock moving at speed vsh, as shown above. Similar to our derivation of the
conservation equation consider the change in mass as the shock moves in a time dt. Just as
before, when deriving a differential equation, the flux in – flux out = rate of change of mass:
( − )= ( − )
(19.18)
( − ) Δ
= =
( − ) Δ
(19.19)
again assuming incompressible flow (constant density and porosity). In dimensionless form:
Δ
=
Δ
(19.20)
This is the difference form of the governing equation for speed. Notice that if there is no
shock, but a smooth change in wavespeed, the speed reduces to Eq. (19.17). Indeed, this is a
more elegant and rapid derivation to find the wavespeed than the cumbersome derivation of
a partial differential equation for volume conservation. In the next section I show how to find
the correct shock and its speed using a graphical construction.
The solution for saturation must represent a monotonic decrease in water saturation from 1-
Sor (fw=1) for vD=0 to Swc (fw=0) for large vD (large distance for a given time). The solution can
be a constant saturation, a smooth variation (called a rarefaction) obeying Eq. (19.13) or a
shock which obeys Eq. (19.20). Mathematically there are many ways to do this, but only one
that makes physical sense – that is, a solution which is the correct physical limit when capillary
pressure (which smooths out the shock) becomes small. It is possible to find this physically
correct solution graphically, as shown in the figure overleaf. This is the Welge construction: a
line is drawn from the initial condition (Sw=Swc; fw=0) which is tangent to the fractional flow
curve. The shock is from the initial condition (this is the right state of the shock) to the
saturation (and fractional flow) where the line hits the fractional flow curve (the left state).
The slope of the line is the change in fractional flow divided by the change in saturation, and
so represents the dimensionless shock speed. Mathematically this can be written as:
∂ Δ
= =
∂ Δ
(19.21)
0.9
0.8
0.7
0.6
Fractional flow
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Water saturation
A diagram illustrating the Welge construction to find the correct shock for the Buckley-
Leverett solution. The sloping dashed line is tangent from the initial conditions to the
fractional flow curve. Where this line hits the fractional flow curve represents the left
state of the shock (the saturation indicated by the dashed vertical line) – the right state
is the initial water saturation (0.2 in this case). The slope of the tangent is the
dimensionless shock speed.
The solution for water saturation as a function of dimensionless speed is then as shown
schematically above for our example case. The smooth part of the profile – the rarefaction –
is found by computing the slope of the fractional flow curves for saturations higher than the
shock front, as a function of saturation. This can be done analytically (from a closed-form
expression of the fractional flow) or – with care – graphically.
Buckley-Leverett solution
0.7
0.6
0.5
Water saturation
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Dimensionless velocity
The solution shown in the figure above, obeys conservation of volume and the boundary
conditions. Other possible shocks either give unphysical solutions (multiple values of
saturation for a given speed) or are unstable physically. The correct shock is self-sharpening.
To explain this, consider the real situation where there is some capillary pressure. The
capillary pressure has a diffusive effect and tends to smear out the shock. At the leading edge
– small saturations – the wavepseed is slower than the shock speed, so this saturation is
caught up by the saturation behind it. To the left side of the shock, any smearing appears to
slow down the saturation. But here, the natural wavespeed will be higher than the shock
speed and so this water will speed up. The net result is a shock which is stable against
perturbations due to capillary pressure. Any other possible solution (and to avoid confusion I
will not present them here) will lead to a shock that will decompose and rearrange as the
stable shock we have just described.
This solution is called the Buckley-Leverett solutions after the authors who first presented it;
the conservation equation (18.13) and its variants are often called the Buckley-Leverett
equation.
Remember that the wave speed IS NOT the same as the particle (tracer) velocity:
(19.21)
The particle speed is the speed with which a single molecule of water moves through the pore
space. This can be explained simply by considering conservation of volume in a system where
the saturation is constant (or over a length where the change in saturation is small). Imagine
that, say, blue water is injected to displace red water. Then is we inject a volume qw=fwqt per
unit area per unit time, this blue water will fill a volume fwqt/(Sw) of the porous medium – in
dimensionless form this is the speed given by Eq. (19.21).
This is distinct from the wavespeed. This is the speed with which a given saturation moves,
not a single particle. Why is this different? One way to consider this is through analogy with
traffic flow. Travelling in a car on a motorway, you are always moving forwards, or stationary.
But now imagine yourself in a helicopter flying above the motorway (and, for a more graphic
illustration this is the M25 during the rush hour, so there are many traffic jams). You can see
waves of traffic density – this is analogous to saturation. If there is an accident, for instance,
the cars at the crash site are stationary – the particle speed is zero. However, a wave of
stationary cars moves backwards up the motorway: the wavespeed associated with a dense
packing of (unmoving) cars is negative.24 So, here, the wavespeed for saturation is a speed
that a particular value of saturation moves, but this is not the same as the speed of a particular
water molecule (the particle speed).
Now we briefly present some definitions useful for understanding the terminology used to
describe these solutions. A spreading wave or rarefaction is when the wave becomes more
diffuse with time – smooth changes in saturation.
A sharpening wave is when the wave becomes less diffuse and sharpens to form a shock.
24
This is why driving when there is an accident (particularly in poor visibility) is so dangerous.
You might think that you have sufficient time (and space) to stop before the car in front comes
to a halt, but in reality a wave of stationary traffic is moving towards you – this is less safe
than there being a brick wall just out of sight.
Indifferent is the case when the wave neither spreads nor sharpens.
A constant state is a fixed saturation with distance (or velocity). This is also an acceptable
solution to the equations.
We can also consider the impact of gravity on our analytical solutions, as shown below with
some schematic curves. If the water flow is uphill, then the water flow is held back, resulting
in a higher shock saturation moving more slowly than a corresponding case with no gravity.
Even if the fractional flow becomes negative, the shock jumps across this region and so we do
not see explicit counter-current flow for water injected at the bottom of the formation.
Note that it is possible that the wavespeed at 1-Sor is not zero (as in the previous examples)
but finite – this is the derivative of the fractional flow at Sw=1- Sor. In this case we have a
constant state from vD=0, to the computed value. As mentioned previously, this is a perfectly
acceptable solution of the conservation equation.
Sw
1-Sor
Gravity
No gravity
Swi
0
0 vD
The Buckley-Leverett solution showing the effect of gravity for water flowing uphill:
buoyancy here leads to a higher slower-moving water shock.
If we flow downhill, then water moves faster and we see a faster-moving, shallower water
shock. If the water fractional flow at vD=0 is 1, then again we cannot observe strictly counter-
current flow. However, it is possible to construct solutions where we have a backwards
moving shown or rarefaction, representing the portion of the fractional flow curve which is
greater than 1.
Sw
1-Sor No gravity
Gravity
Swi
0
0 vD
The Buckley-Leverett solution showing the effect of gravity for water flowing downhill:
buoyancy in this case leads to a lower faster-moving water shock.
The final stage in the analysis is to use the analytical solution for saturation as a function of
speed (and hence, for a given time, as a function of distance) to compute recovery. In our
previous discussions we have always discussed how recovery is a function of how much water
is injected (the pore volumes of water injected). The way to derive at a recovery calculation
is first to consider the average saturation. The recovery is proportional to the change from
the initial (usually connate) water saturation to this average value. Specifically, the average
saturation in a domain after breakthrough – that is where the shock front has already reached
the production well. Consider the schematic saturation profile shown on the next page.
Sw
1-Sor
Sw1
Swi
0
0 xD1 xD
The Buckley-Leverett solution as a function of dimensionless distance employed to
compute the average saturation and hence recovery. xD1 represents the production well
(xD=1) and we compute the average saturation between this and the injector at xD=0.
The average saturation behind xD1 is easy to define, but the mathematics is somewhat tedious
to produce a tractable solution:
1
̅ ( )=
(19.22)
Integrate by parts:
1
̅ ( )= −
(19.23)
df w
We can write xD
1
t D , and thus Eq. (19.23) becomes:
dSw S w S 1w
1−
̅ ( )= − = − = + (1 − )= +
(19.24)
Notice the prime on the final equation, denoting the derivative. Graphically this can be
represented as the extension of the tangent of the curve at some location on the rarefaction
to fw=1 gives the average saturation behind the front. This construction is shown
schematically below. From a practical perspective, choose any saturation above the shock-
front value, draw a tangent and find the saturation value when this tangent reaches fw=1.
1
fw 1
fw S w1
0
0 Swc Sw Sw 1-Sor
A schematic of the construction used to find the average saturation when the saturation
at the well (xD=1) is some arbitrary value Sw1.
We now use this construction to compute the recovery – pore volumes of oil produced – as a
function of time, or pore volumes of water injected: NpD vs. tD. We know that by definition,
the average saturation = Swi +NpD, where NpD are the pore volumes of oil produced.
Before breakthrough, the reservoir volume of oil produced is equal to the reservoir volume of
water injected and so NpD=tD. Breakthrough occurs when the shock moving at dimensionless
speed vshD reaches xD=1: this is a dimensionless time tD=1/vshD. So, the first step in constructing
a recovery curve is to draw a straight line of unit slope from NpD=tD =0 to NpD=tD =1/vshD.
After breakthrough, the recovery curve has a slope less than one, indicating the production of
both oil and water. What is the maximum recovery? This is 1-Sor-Swc: you cannot produce
more oil than this. When will this occur? If the wavespeed for Sw=1-Sor is zero, this happens
at infinite time, so the maximum recovery is met asymptotically at infinite time. If the
wavespeed has a finite minimum values, say vDmin, then maximum recovery is met at a time
1/vDmin. Now construct one or two points in between. Choose a value of saturation in the
rarefaction. Find the tangent through this saturation value on the fractional flow curve. The
slope of the tangent is the wavespeed, vD, while the intersect when fw=1 is the average
saturation. The recovery is the average saturation minus the initial value, while the
dimensionless time tD=1/vD.
The resultant plot for our example case is shown in the figure below. In words, the procedure
does seem a little bewildering – the only way to learn this is through performing the exercise
yourself.
1. Given relative permeability curves, you first compute the fractional flow and plot this
as a function of saturation.
2. Perform the Welge construction to find the shock front saturation and shock speed.
4. Plot the saturation as a function of dimensionless velocity: in the rarefaction the
speed is the slope of the fractional flow curve.
5. Compute the dimensionless recovery (pore volumes produced) as a function of
dimensionless time (pore volumes injected). Before breakthrough, since we have
assumed that the oil and water are incompressible, these two quantities are the same.
After breakthrough of water (when the shock front reaches the production well), the
pore volumes of oil produced are less than the pore volumes of water injected, since
some water is produced as well. You find the recovery from choosing points in the
rarefaction and extrapolating a tangent on the fractional flow curve to fw=1, as
described above.
Dimensionless recovery
0.5
0.45
0.4
Pore volumes of oil produced
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Pore volumes of water injected
Pore volumes produced as a function of pore volumes of water injected for our example
case. The red line has unit slope and occurs before breakthrough. After breakthrough
the black line was constructed using the methodology described in the text. The
ultimate recovery is 0.5, but this is only reached asymptotically once an infinite number
of pore voluems are produced. Rarely is more than pore volume injected in a reservoir
waterflood and so there is no point showing recovery beyond, say, 2 pore volumes, as
here.
This is an important exercise, and is performed by any reservoir engineer to assess waterflood
recovery. In the end the behaviour at the field scale is determined by both this analysis and
the geology of the field, which defines the preferential flow paths for the injected water, and
the placement of the wells – an assessment of this usually requires reservoir simulation
methods and lies outside the scope of these notes.
We will now discuss different types of behaviour physically and the implications for recovery.
Note though, that this rigorous analysis can substitute for the rather empirical approach used
previously and should be used to assess recovery when both the relative permeabilities and
viscosities are known. There are generically three types of solution.
1. Classic Buckley-Leverett. This is the case we have shown with a rarefaction and
a shock. This is observed for most power-law relative permeabilities and
The last point to note that a Buckley-Leverett style analysis is not confined to a one-
dimensional analytical analysis: it is also useful to compare field-scale recovery (either real
data or simulation predictions) to Buckley-Leverett predictions. It is straightforward (with a
very careful scrutiny of the definitions) to convert data for oil produced at surface conditions
into pore volumes produced, and real time into pore volumes of water injected. Since the real
reservoir is heterogeneous, the recovery will always lie below that from an idealized Buckley-
Leverett analysis, but does serve as a very useful basis of comparison. This is also helpful in
field management: the closer the predicted (or real) recovery to Buckley-Leverett, the better
the injection wells are contacting and sweeping the reservoir. It also serves as a guide to how
much water needs to be injected to achieve optimal recovery.
We can now return to our previous analysis of relative permeability and use these curves to
compute recovery for a linear displacement. It is important to emphasize that in most field
cases we rarely inject more than one pore volume of water. Hence, the fact that very low oil
saturations can be achieved by oil layer drainage when hundreds or thousands of pore
volumes are injected is not economically relevant, even if it can be seen in laboratory
experiments. Of much greater import is the shock front saturation: this generally determines
the local efficiency of waterfooding, which a high shock front being most favourable.
In general, strongly water-wet samples give a Buckley-Leverett displacement that is all shock,
with the maximum recovery reached at breakthrough. This is efficient, but the high residual
saturation makes the process less than ideal. Strongly oil-wet systems also give poor recovery
– often even poorer recovery – since the oil breaks through very early and there is only
substantial production after breakthrough, where water is also produced. This is economically
unfavourable. The best recovery is for wettability states in between: mixed-wet or weakly
water-wet rocks. One of the new research areas at the moment is in how to control
wettability through adjusting the chemistry of the injected brine. This is the idea behind low
salinity waterflooding that aims to adjust the rock wettability to a favourable near water-wet
state.
The reality is that for any given system, a careful assessment of relative permeability is
warranted. First I show below some results from Blunt (1998) from pore-scale network
modelling. Here the relative permeabilities are computed for different fractions of oil-wet
pores, the fractional flow curves determined and from this recovery as a function of pore
volumes injected is found. This is exactly the exercise – bar finding the relative permeabilities
in the first place – that I will expect you to be able to do by hand, and which has been described
above.
Note in this case the trend of residual oil saturation with wettability, showing the lowest
residual – thanks to layer drainage – for the most oil-wet case. The wettability indices for oil
and water are shown, indicating mixed-wet behaviour. For around 1 pore volume of water
injected, it is not the oil-wet case, but the weakly water-wet system (with no oil-wet pores),
that gives the best recovery. Adding some oil-wet pores reduces recovery, since these pores
are trapped during waterflooding and do not themselves connect through the system,
meaning that layer drainage cannot allow low residual saturations to be reached. It is only for
higher oil-wet fractions, when these pores connect across the system, that low residual
saturations are seen. The completely oil-wet case only gives the highest recovery after the
injection of around 100 pore volumes of water – something that never practically occurs in a
field setting.
Residual oil saturation, wettability indices and oil recovery from network modelling of a
mixed-wet system, with a faction f of oil-wet pores indicated. Taken from Blunt (1998).
I will now illustrate the remarks above with some experimental data on sandstones. Note,
however, that for the carbonates discussed earlier, the most favourable conditions for
waterflooding were more mixed-to-oil-wet than shown here.
The example I choose uses the computations of relative permeability for mixed-wet Berea
sandstone made by Valvatne and Blunt (2004). In this work, a fixed fraction of the pores
contacted by oil after primary drainage became oil-wet. Different initial saturations of water
(that is saturations after primary drainage) were considered: if the initial water saturation is
high, the system appears to be more water-wet, as the pores that remain full of water remain
water-wet. As the initial saturation decreases, the system becomes more oil-wet. This is the
trend with height that is observed in the transition zone of an oil reservoir.
The graphs below show measured recovery profiles for waterflooding from Jadhunandan and
Morrow (1995) – with different initial water saturations – compared to different predictions
using pore-scale modelling. For this sandstone sample, we see unfavourable recovery for the
most water-wet case (because of the high residual saturation) and for the most oil-wet case
(because of early water breakthrough and the slow drainage of oil layers). The ideal cases
have an intermediate initial water saturation and are overall weakly water-wet to mixed-wet
in character, meaning that the Amott-Harvey wettability index is close to zero.
Measured (left) and predicted (middle and right) oil recoveries as a function of
pore volumes injected in a mixed-wet Berea sandstone. The different initial
water saturations are indicated. Note that the most favourable recoveries occur
for intermediate saturations. The two modelling predictions use slightly
different assignments of wettability.
This behaviour, as alluded to previously, is different from the trend we expect in carbonates,
where the most favourable recoveries are seen when most of the pores are oil-wet: the
difference is a result of the connectivity of the pore space, the nature of the pore and throat
size distributions and the local variations in wettability (contact angle). Carbonates display a
very wide range of connectivity and pore sizes, and can have very low recoveries in the water-
wet limit; however, a definitive characterization requires detailed analysis of the sample of
interest coupled with an accurate assessment of wettability. At present, we do not have a
way to assign contact angle on a pore-by-pore basis unambiguously, and so, at present, we
have to rely on macroscopic measurements of wettability (such as Amott index) to tune the
contact angle distribution in our pore-scale models.
One last comment: the pore volumes produced is NOT the recovery factor shown in the graphs
above. The recovery factor, RF, is a ratio of the volume of oil produced to the total of oil
initially in place. It is easy to relate these two quantities as follows:
=
(1 − )
(19.25)
We will now present a solution for spontaneous imbibition, where displacement is controlled
entirely by capillary forces. This is a useful complement to the Buckley-Leverett solution. It is
valuable experimentally as a way to determine, or at least constrain, capillary pressure and
relative permeability. It is also useful for the analysis of recovery in fractured reservoirs, as
discussed previously.
The formulation here is quite new in the literature. While the solutions were first proposed
by McWhorter and Sunada (1990), it was not until the work of Schmid and Geiger (2012) that
it was appreciated that this was indeed a closed-form solution generally applicable for
spontaneous imbibition.
Before wading into the mathematical details, let us review the physical situation. It is
illustrated below, where bubbles of non-wetting phase escape from a core when the wetting
phase imbibes from all sides. It is similar to the bubbles of air seen around a cube of sugar
dropped into a drink – if we ignore the dissolution.
Photograph illustrating imbibition in a rock core – the bubbles are the displaced non-
wetting phase.
For clarity, we start with the conservation Eqs. (18.3) and (18.2) which are written here in
terms of the water Darcy velocity (rather than the total velocity, since this will be zero):
+ =0
(20.1)
= + +( − )
(20.2)
For spontaneous imbibition, we ignore gravitational forces (assume that they are either small
compared to capillary forces at the core – cm – scale, or that the displacement is horizontal)
and the total velocity. Setting the total velocity qt to zero means that no fluid is injected and
the flow is counter-current: the movement of water into the porous medium is matched
exactly by the volume of oil (or gas) that leaves. Then we write:
(20.3)
+ =0
(20.4)
Assuming a constant porosity we can write Eq. (20.4) as a non-linear diffusion equation:
= ( )
(20.5)
( )=−
(20.6)
Note the negative sign. D is positive, so we assert that the gradient of the capillary pressure
as a function of water saturation is always negative. You should now appreciate that this is
indeed correct.
The boundary conditions are a porous medium containing initially irreducible water (S=Swc)
and a non-wetting phase (which we call oil for convenience here). At the inlet – x=0, we
maintain a capillary pressure of zero. In a strongly water-wet system, the saturation will be
1-Sor; in any other case this will simply be the saturation at which the capillary pressure is zero,
which we define as S*.
=
√
(20.7)
where we assume that the solution can be stated as Sw() only. I also will state that we can
write:
(20.8)
for some capillary fractional flow F(Sw). We assume that F has a maximum value F*=F(S*) and
is zero for the irreducible water saturation: F(Swc)=0.
Note the approach here. I have assumed a certain functional form of the solutions, based –
please note – on both the previous solution for diffusion and the Buckley-Leverett analysis.
However, this is an educated guess and we have to test if we can solve the governing partial
differential equations and the boundary conditions. There is no a priori manner to know –
before we start – if this approach is correct. Despite what you may be told in mathematics
classes, solving partial differential equations is more inspiration (guessing the correct
approach) than application (of all the methods you are given in class).
=−
2
(20.9)
1
=
√
(20.10)
+2 =0
(20.11)
We integrate once:
= −2
(20.12)
where the integration constant is zero since we define F(Swc)=0 and also D(Swc)=0. Then
substitute in F from Eq. (20.8) to find:
= −2
(20.13)
Eq. (20.13) is the key equation to define F and hence construct a solution.
In a formal mathematical sense, the solution can be expressed in closed form simply by
integrating Eq. (20.13) twice – this is the solution presented in Schmid and Geiger. The
problem – that appears – at first sight – somewhat off-putting is that the solution is expressed
in terms of implicit integrals: that is the integral to find F involves F itself. Indeed, it is
presented in terms of two implicit integrals, since we have to define F(1-Sor), noting that F is a
dimensional quantity and so we cannot set it to 1.
We have already defined one boundary condition on F: F(Swc)=0. Since we have a second-
order equation for F, we require two conditions. The second, applicable in most cases, is that
for Sw=1-Sor, dF/dSw=0 (the saturation front does not move at the inlet, which again makes
physical sense).
However, before proceeding, let is first study the implications for the amount of water that
enters the porous medium. The water Darcy velocity can be found from Eqs. (20.3) and (20.6):
=−
(20.14)
=−
√
(20.15)
′′
= =
2√ 2√
(20.16)
The inlet flux is then related to the value of F*. The total amount of water that enters the
system, Qw is:
∗
∗
= = √
2√
(20.17)
The amount imbibes scales as the square-root of time. Note that this expression differs from
the solutions presented previously, where we allowed the flux injected to reach a maximum
in a block of finite size: here we consider only the early-time behaviour before the imbibing
front has reached a boundary.
We will find a solution using a simple numerical approach. For given relative permeabilities
and capillary pressures, we compute D. We then solve Eq. (20.13) numerically in a
spreadhseet starting from S* and then decreasing saturation in small increments. We guess
F* and impose F’(S*)=0. We then iterate to find the value of F* such that when Sw=Swc, F=0. I
will not go through the details, but it is readily computed using a backwards difference
scheme. I will show some example results below. Notice that using a Buckley-Leverett
analogy, this is a simple case, since there are no shocks: the saturation profile is smeared out
and we have, technically, an all rarefaction solution.
The capillary pressure used in my example calculation for imbibition. This resembles a
primary drainage curve, but in fact – as we see below – this is a mixed-wet case. I
assume that for saturations beyond S*=0.6 the capillary pressure is negative.
I show the capillary pressure and relative permeabilities in the saturation range for which
there is spontaneous imbibition. I then show the dimensionless fractional flow that varies
between 0 and 1: this is defined as f=F/F*. Similarly, I can define dimensionless wavespeeds
D=/F*. In the example I show F*=2.284×10-4 m/√s. I then show the saturation as a function
of : this is the full solution, with the distance moved by the saturation scaling not linearly
with time (as for water injection in the Buckley-Leverett analysis) but as the square root of
time.
The relative permeabilities used in my example calculation for imbibition. The values
are truncated when the capillary pressure is zero.
The dimensionless wavespeed found from the derivative of the fractional flow shown
above.
The same curve as before, but now multiled by F* to give the dimensional wavespeed.
If, instead, we had experimental measurements, then we can use the measured saturation
profiles, obtained from in situ scanning (which give us F’) to find D using Eq. (20.13). This is
now the topic of on-going research. My hope is that combining pore-scale modelling from
images, macroscopic corefloods using Buckley-Leverett theory, this imbibition solution and
steady-state measurements of relative permeability, we can readily and very reliably
determine capillary pressure and relative permeability. At present, the subject woefully lacks
good experimental data; the methods described in these notes offer a new opportunity to
produce results that are reliable and accurate.
A complete discussion of this is beyond the scope of these notes, but uses the same ideas –
derive an appropriate conservation equation and identify wave and shock speeds.
Needless to say, however, this is a powerful and relatively simple way to assess recovery and
displacement in porous media and serves as a useful complement to more sophisticated
numerical approaches. In any event, numerical models require relative permeability curves
as input and it is important to be able to understand their impact on the flow behaviour.
In the end, to describe flow in heterogeneous reservoirs with many wells and complex
constraints on pressure and rates, it is necessary to perform a numerical analysis, solving the
flow and transport equations presented in these notes in three dimensions. This is a rich and
fascinating topic in its own right. However, it is important to retain a physical insight into the
displacement and recovery processes and how core-scale analysis and measurements relate
to field-scale recovery.
Here are the references of the papers mentioned in the notes as well as other papers that are
useful for reference and further reading.
Many of these papers can be accessed through the Society of Petroleum Engineers database,
if they are not available through Web of Science or Google Scholar – www.onepetro.org.
M Akbarabadi and M Piri “Relative permeability hysteresis and capillary trapping characteristics of
supercritical CO2/brine systems: an experimental study at reservoir conditions,” Advances in Water
Resources, 52, 190–206 (2013).
W G Anderson, “Wettability literature survey part 1: Rock/oil/brine interactions and the effects of
core handling on wettability, J. Petrol. Technol., 38(10), 1125–1144, doi: 10.2118/13932-PA. (1986).
W G Anderson, “Wettability literature survey-part 6: the effects of wettability on waterflooding,” J.
Petrol. Technol., 39(12), 1605–1622, doi:10.2118/16471-PA. (1987).
C H Arns, M A Knackstedt, V Pinczewski and W B Lindquist, “Accurate estimation of transport
properties from microtomographic images,” Geophysical Research Letters 17, 3361-3364 (2001).
C H Arns, F Bauget, A Limaye, A Sakellariou, T J Senden, A P Sheppard, R M Sok, V W Pinczewski, S
Bakke, L I Berge and P-E Øren, “Pore-scale characterization of carbonates using X-ray
microtomography,” SPE Journal 10(4) 475-484 (2005).
J Y Arns, A P Sheppard, C H Arns, M A Knackstedt, A Yelkhovsky and W V Pinczewski, “Pore-level
validation of representative pore networks obtained from micro-CT images,” Proceedings of the
Annual Symposium of the society of Core Analysis, SCA2007-A26, Calgary, Canada (2007).
J S Aronofsky, L Masse, and S G Natanson, “A model for the mechanism of oil recovery from the
porous matrix due to water invasion in fractured reservoirs,” Petrol. Trans. AIME, 213, 17–19 (1958).
S Bakke and P-E Øren, “3-D pore-scale modelling of sandstones and flow simulations in the pore
networks.” SPE Journal 2 136-149 (1997).
F E Bartell and H J Osterhof, Ind. Eng. Chem. 19, 1277 (1927).
S E Buckley and M C Leverett, “Mechanisms of fluid displacement in sands,” Trans. AIME, 146, 107116
(1942).
H S Carslaw and J C Jaeger, Conduction of Heat in Solids, 2nd Edition, Oxford Science Publications,
Clarendon Press, Oxford (1946).
I Chatzis and N Morrow, “Correlation of Capillary number relationships for sandstone,” SPE Journal,
24(5), 555–562, doi:10.2118/10114-PA (1984).
J H Dunsmuir, S R Ferguson, K L D'Amico and J P Stokes, “X-ray microtomography. A new tool for the
characterization of porous media,” SPE 22860, Proceedings of the 1991 SPE Annual Technical
Conference and Exhibition, October 1991, Dallas (1991).
I Fatt, “The network model of porous media I. Capillary pressure characteristics,” Trans AIME 207
144-159 (1956).
T M Okasha, J J Funk, and H N Rashidi, “Fifty years of wettability measurements in the Arab-D
carbonate reservoir, paper SPE 105114. Proceedings of the SPE Middle East Oil and Gas Show and
Conference, 11–14 March 2007, Manama, Kingdom of Bahrain (2007).
T W Patzek, “Verification of a complete pore network simulator of drainage and imbibition,” SPE
Journal, 6 144-156 (2001).
M Sahimi, “Flow and transport in porous media and fractured rock,” Wiley-VCH Verlag GmbH. (1995).
R A Salathiel, “Oil recovery by surface film drainage in mixed-wettability rocks,” SPE Journal, 25(10),
1216–1224, doi:10.2118/4104-PA (1973).
K S Schmid and S Geiger, “Universal scaling of spontaneous imbibition for water-wet systems,” Water
Resources Research, 48, W03507, doi:10.1029/2011WR011566, (2012).
N C Wardlaw and R P Taylor, “Mercury capillary pressure curves and the intepretation of pore
structure and capillary behaviour in reservoir rocks,” Bulletin of Canadian Petroleum Geology, 24(2),
225–262 (1976).
D Wilkinson and J F Willemsen, “Invasion percolation: a new form of percolation theory,” J Phys A, 16,
3365–3376 (1983).
X Zhou, N R Morrow and S Ma, “Interrelationship of wettability, initial water saturation, aging time,
and oil recovery by spontaneous imbibition and waterflooding,” SPE Journal, 5(2), 199–207,
doi:10.2118/62507-PA (2000).
These are presented in date order. Most of these papers can be downloaded from our
website:
http://www.imperial.ac.uk/earth-science/research/research-groups/perm/research/pore-
scale-modelling/
M J Blunt, "Physically Based Network Modeling of Multiphase Flow in Intermediate-Wet Media" Journal
of Petroleum Science and Engineering, 20 117-125 June (1998).
P H Valvatne and M J Blunt, “Predictive pore-scale modeling of two-phase flow in mixed wet media,”
Water Resources Research, 40, W07406, doi:10.1029/2003WR002627 (2004).
H Okabe and M J Blunt, “Prediction of permeability for porous media reconstructed using multiple-
point statistics,” Physical Review E 70, 066135 (2004).
M S Al-Gharbi and M J Blunt, “Dynamic network modeling of two-phase drainage in porous media,”
Physical Review E 71, 016308 (2005).
M Piri and M J Blunt, “Three-dimensional mixed-wet random pore-scale network modeling of two- and
three-phase flow in porous media. I. Model description,” Physical Review E 71, 026301 (2005).
M Piri and M J Blunt, “Three-dimensional mixed-wet random pore-scale network modeling of two- and
three-phase flow in porous media. II. Results,” Physical Review E 71, 026302 (2005).
P H Valvatne, M Piri, X Lopez and M J Blunt, “Predictive Pore-Scale Modeling of Single and Multiphase
Flow,” Transport in Porous Media 58, 23–41, doi:10.1007/s11242-004-5468-2 (2005).
Z Tavassoli, R W Zimmerman and M J Blunt, “Analytic Analysis for Oil Recovery During Counter-Current
Imbibition in Strongly Water-Wet Systems,” Transport in Porous Media 58, 173–189,
doi:10.1007/s11242-004-5474-4 (2005).
B Amaechi, S Iglauer, C H Pentland, B Bijeljic and and M J Blunt, “An Experimental Study of Three-Phase
Trapping in Sand Packs,” Transport in Porous Media 103(3) 421-436 (2014).
A Q Raeini, B Bijeljic, M J Blunt, “Modelling capillary trapping using finite-volume simulation of two-
phase flow directly on micro-CT images,” Advances in Water Resources, 83 102–110 (2015).
J P Pereira Nunes, M J Blunt and B Bijeljic, “Pore-scale simulation of carbonate dissolution in micro-CT
images,” J. Geophys. Res. Solid Earth, 121, 558–576 (2016).
H P Menke, M G Andrew, M J Blunt and B Bijeljic, “Reservoir condition imaging of reactive transport in
heterogeneous carbonates using fast synchrotron tomography — Effect of initial pore structure and
flow conditions,” Chemical Geology, 428 15-26 (2016).
K Singh, B Bijeljic, B and M J Blunt, “Imaging of oil layers, curvature and contact angle in a mixed-wet
and a water-wet carbonate rock,” Water Resources Research, 52(3), 1716-1728 (2016).
N Alayafei and M J Blunt, “The effect of wettability on capillary trapping in carbonates,” Advances in
Water Resources, 90, 36-50 (2016).
A S Al-Menhali, H P Menke, M J Blunt and S C Krevor, “Pore Scale Observations of Trapped CO 2 in Mixed-
Wet Carbonate Rock: Applications to Storage in Oil Fields,” Environmental Science and Technology,
50(18), 10282–10290 (2016).
Z Alhashmi, M J Blunt and B Bijeljic, “The Impact of Pore Structure Heterogeneity, Transport, and
Reaction Conditions on Fluid–Fluid Reaction Rate Studied on Images of Pore Space,” Transport in
Porous Media, 115, 215–237 (2016).
A I Bajwa and M J Blunt, “Early-Time 1D Analysis of Shale-Oil and -Gas Flow,” SPE Journal, 21(4), 1254—
1262 (2016).
Q Lin, Y Al-Khulaifi, M J Blunt and B Bijeljic, “Quantification of sub-resolution porosity in carbonate rocks
by applying high-salinity contrast brine using X-ray microtomography differential imaging,” Advances
in Water Resources, 96, 306–22 (2016).
T Saif, Q Lin, K Singh, B Bijeljic and M J Blunt, Dynamic imaging of oil shale pyrolysis using synchrotron
X-ray microtomography,” Geophysical Research Letters, 43, 6799–6807 (2016).
J P Pereira Nunes, B Bijeljic and M J Blunt, “Pore-space structure and average dissolution rates: A
simulation study, Water Resources Research, 52, 7198–7212 (2016).
H Nooruddin and M J Blunt, “Analytical and numerical investigations of spontaneous
imbibition in porous media,” Water Resources Research, 52, 7284–7310 (2016).
L Leu, A Georgiadis, M J Blunt, A Busch, P Bertier, K Schweinar, M Liebi, A Menzel, and H Ott, “Multiscale
Description of Shale Pore Systems by Scanning SAXS and WAXS Microscopy,” Energy Fuels, 30(12),
10282–10297 (2016).
N Alyafei, A Al-Menhali and M J Blunt, “Experimental and Analytical Investigation of Spontaneous
Imbibition in Water-Wet Carbonates,” Transport in Porous Media, 115, 189-207 (2016).
K S Schmid, N Alyafei, S Geiger and M J Blunt, “Analytical Solutions for Spontaneous Imbibition:
Fractional-Flow Theory and Experimental Analysis,” SPE Journal, 21(6), 2308-2316 (2016).
W-B Bartels, M Rücker, S Berg, H Mahani, A Georgiadis, A Fadili, N Brussee, A Coorn2, H van der Linde,
C Hinz, A Jacob, C Wagner, S Henkel, F Enzmann, A Bonnin, M Stampanoni, H Ott, M Blunt and S M
Hassanizadeh, “Fast X-Ray Micro-CT Study of the Impact of Brine Salinity on the Pore-Scale Fluid
Distribution During Waterflooding,” Petrophysics, 58(1), 36-47 (2017).
These problems are useful to understand much of the material in the notes.
1. Define in a single sentence the following terms: (i) capillary pressure; (ii) drainage; (iii)
imbibition; (iv) forced water injection.
2. Draw a typical oil/water capillary pressure curves for a water-wet sandstone with a
porosity of 0.2 and a permeability of 100mD. Draw primary drainage, imbibition and
secondary drainage curves. Also indicate on the graph typical values for the residual
oil saturation and connate water saturation and typical values for the capillary
pressure. These need only be estimates, but please explain clearly why you chose
these values. Also explain why the primary and secondary drainage curves are
different and why the imbibition curve is lower than the secondary drainage curve.
3. You are planning a surfactant flood. You perform some core flood experiments to find
the effect of capillary number on the residual oil saturation. You find a correlation of
the form:
N cap
Sor Max0,0.4
0.08
You now perform a reservoir-scale flood. The injection and production wells are 100
m apart and have a pressure drop of 10 atm between them. The rock permeability is
200 mD. With surfactant the oil/water interfacial tension is 0.01 mN/m. What is the
predicted residual oil saturation? What pressure difference between the wells is
necessary to remove all the oil?
f (r )dr 1
0
All the tubes are aligned horizontally and have the same pressure drop across them.
They are all of the same length. The flow (volume per unit time) in each tube is given
by:
r4
Q P
8 l
The total volume of each tube = r2L. The tubes are water-wet and are initially filled
with water. Oil is now injected into the tubes. What size tubes does the oil
preferentially occupy? The last tube to be filled with oil has a radius R. Derive an
expression for the oil saturation and the oil relative permeability.
5. You are planning a tracer test in a portion of a reservoir that is entirely full of water.
The tracer dissolves in water and flows with the water. You also know that the tracer
absorbs to the reservoir rock. The mass of tracer that absorbs per unit volume of the
reservoir, a is given by the following equation:
a ac
derive a conservation equation for tracer in one dimension, where q is the water
volumetric flux per unit area and is the rock porosity. You may assume that q and
are both constant.
With what speed does the tracer flow through the rock? Find this speed when q = 10-
6
ms-1, = 0.2 and a = 3.
c
flux is FD D , where D is the diffusion coefficient. Since the tracer is radioactive
x
its concentration in a static fluid with no concentration gradient will decrease with
time as c(t ) c(t 0)et , which is the same as saying that with no flow c c
t
7. Unit conversions
A committee is set up to increase North Sea oil production by 1,000,000 barrels per
day. The committee members disagree on what units to use. For all the sections
below express 1,000,000 barrels per day in the unit systems described.
(a) Milton Keynes, an economist, wants to use millions of $ per year. The oil price is $105
per barrel.
(c) Abdus Goldstein, a theoretical physicist, wants to use units in which h/2 = 1, c = 1
and G = 1. h/2 = 1.055x10-36 Js, c = 3.0x108 ms-1, G = 6.7x10-11 m3kg-1s-2.
(d) Jerry R. Beltbuckle III, an oil industry representative, prefers to use acre-feet per
month. What is the target in (i) February 1999, (ii) December?
f
v
S
r 2h
v
Qt
where r is the radial distance from the well, h is the perforated interval, t is the time ,
is the porosity and Q is the flow rate at the well.
10. For each of the examples below provide a graph of: (i) Sw as a function of vD = xD/tD;
and (ii) pore volumes of oil produced NpD as a function of tD.
k rw k rw
max S w S wc a
1 S or S wc a
k ro k romax
1 S or S w b
1 S or S wc b
o k rw
max
M
w k romax
max max
(a) A strongly water-wet rock with a = 3, b = 1, k rw = 0.18, k ro =0.9, Swc = 0.2, Sor = 0.4
and w = o (M = 0.2).
max max
(b) An oil-wet rock with a = 1, b = 3, k rw = 0.9, k ro =0.18, Swc = 0.2, Sor = 0.1 and w =
o (M = 5).
Hints: Yes, this is a tedious exercise, but after it you will really know how to perform
a Buckley-Leverett analysis. Either plot out the fractional flow and find the shock
height graphically by hand, or do everything analytically/numerically. It is not possible
to find a closed form expression for the shock saturation for this type of relative
permeability, but you could write a small computer program to find everything you
need automatically. Once you have worked it out for one case, the others should be
easy.
(i) Assuming a simple aquifer model, estimate the aquifer size and the original gas in place. You
may neglect the compressibility of the connate water and rock in the gas field itself. (25
marks)
(ii) Gas sales are planned for this field until the reservoir pressure drops to 2 MPa. Make an
APPROXIMATE estimate of the Z factor at 2 MPa. Think carefully about the limit of Z for low
pressure. (5 marks)
(iii) Estimate the gas produced at 2 MPa and the recovery factor. (6 marks)
(iv) Comment on your answer to part (iii). Is your answer physically reasonable and if not why
not? What will happen that will prevent the reservoir pressure reaching 2 MPa? (8 marks)
(v) What further information would you need to estimate the likely recovery factor for the field
and the pressure of likely abandonment? (6 marks)
2. You have the following data for an oil reservoir. (50 marks total)
Np Gp Rs Bo P Bg
k ro 0.8S o 0.4
2
o = 0.0025 Pa.s
w = 0.001 Pa.s
= 0.3
L = 30 cm
Q = 0.1 cm3 / minute
A = 1 cm2
(i) For the Buckley-Leverett analysis what assumptions are made about the displacement? (6
marks)
(ii) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. (20 marks)
(iii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iv) Find the time to breakthrough of water for this core. (6 marks)
(v) After injecting water for 10 hours, what is the recovery factor and what volume of oil has
been recovered? (8 marks)
UNIVERSITY OF LONDON
PE 300(ii) RESERVOIR ENGINEERING I (RESERVOIR MECHANICS AND SECONDARY RECOVERY)
May 2000
Answer any THREE questions
1. You have the following production data for a dry gas field. (50 marks total)
Pressure Z Gp
(MPa) (108 m3)
25 0.85 0
24 0.86 6.09
23 0.87 11.8
22 0.88 17.1
21 0.89 22.1
2. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(i) Use the material balance equation to determine the type of reservoir drive, the amount of
initial oil in place and the size of the gas cap (if any). You may assume that there is no water
influx and that the compressibility of the formation is negligible. (25 marks)
(ii) Explain carefully the trend in Rp observed for this field. Is the field above or below the bubble
point at 290 atm? As part of your answer explain what is meant by the critical gas saturation.
Why is production normally stopped once the critical gas saturation is reached in the
reservoir? (10 marks)
(iii) What is the gas saturation when the reservoir pressure is 290 atm? The connate water
saturation is 0.25. (5 marks)
(iv) The critical gas saturation is 0.2. Estimate the recovery factor at this saturation. What
parameter do you need to estimate? Make a sensible estimate of this parameter, based on
the other values that have been measured. (10 marks)
k ro 0.8S o 0.3
2
o = 0.004 Pa.s
w = 0.001 Pa.s
= 0.2
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (20 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) Between wells the average total velocity is approximately 0.1 m/day. If the injection and
production wells are 500 m apart, how long will it be before water breaks through? (6 marks)
(iv) What is the reservoir water fractional flow at water breakthough? If Bo = 1.4 and Bw = 0.9,
what is the surface fractional flow at water breakthrough? (8 marks)
(v) After injecting water for 2 years, what pore volumes of oil are recovered? (6 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2001
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the T.H. Huxley School (Engineering)
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
1. You have the following production data for a large, dry gas field that is produced by a strong
natural water drive. (50 marks total)
Pressure Gp
(MPa) (108 scf)
41 0
40 0.571
39 1.123
38 1.658
37 2.175
2. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(108 stb) (scf/stb) (scf/stb) (rb/stb) (atm) (rb/scf)
0 0 800 1.634 280 0.00345
2.33 900 700 1.603 260 0.00387
3.61 950 600 1.584 240 0.00412
4.59 970 500 1.554 220 0.00435
(i) Use the material balance equation to determine the type of reservoir drive, the amount of
initial oil in place and the size of the gas cap (if any). You may assume that there is no water
influx and that the compressibility of the formation is negligible. (25 marks)
(ii) What is the average gas saturation in the reservoir initially occupied by oil when the reservoir
pressure is 220 atm? The connate water saturation is 0.25. Comment on the likely
consequences of having this amount of gas in the reservoir. What could be done to prevent
the production of excessive amounts of gas? (13 marks)
(iii) Re-injection of produced gas into the top of the reservoir is being considered for this field.
Comment on the advantages and disadvantages of such a strategy with reference to your
answer to part (ii). (12 marks)
k ro 0.6S o 0.35
2
o = 0.003 Pa.s
w = 0.001 Pa.s
= 0.2
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (20 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (8 marks)
(iii) The injection and production wells are 300 m apart. The average cross-sectional area of the
reservoir is 1,600 m2. Water is injected at a rate of 120 m3/day (measured at surface
conditions). At the initial reservoir pressure Bo = 1.3 and Bw = 0.96. After 1,000 days, Bo =
1.4 and Bw = 0.96. What is the recovery factor (based on surface volumes)? (12 marks)
(iv) At late times, the reservoir pressure is allowed to fall. Assuming that the Buckley-Leverett
analysis is still correct, find the recovery factor after the injection of 200,000 m3 water (at
surface conditions) when Bo = 1.15 and Bw = 0.96. Comment on your answer compared to
part (iii) How can any apparent inconsistency be resolved? (10 marks)
Write a brief essay that amplifies and explains the remarks above. Discuss each of the seven
advantages listed above and comment on the cases where streamline-based simulation does not
work so well. The essay must be written in your own words.
UNIVERSITY OF LONDON
MSc. EXAMINATION 2002
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
1. You have the following production data for a dry gas field that is produced by a strong natural
water drive. (50 marks total)
Pressure Gp Bg
(MPa) (1010 scf) (rb/scf)
30 0 0.000560
29 4.52 0.000575
28 9.08 0.000595
27 13.55 0.000620
26 17.85 0.000650
The initial water saturation is 0.25.
(i) Define the terms dry gas, wet gas and gas condensate. (7 marks)
(ii) Assuming a simple aquifer model, estimate the original gas in place and the value of the
aquifer size times the aquifer compressibility. You may neglect the compressibility of the
connate water and rock in the gas field itself. (25 marks)
(iii) At a pressure of 24 MPa the wells are watered out and there is no further gas production. If
we assume that at this point, the entire reservoir has been swept by gas, estimate the
residual gas saturation. (18 marks)
2. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(106 stb) (scf/stb) (scf/stb) (rb/stb) (MPa) (rb/scf)
(i) Use the material balance equation to determine the type of reservoir drive, the amount of
initial oil in place and the size of the gas cap (if any). You may assume that there is no water
influx and that the compressibility of the formation is negligible. (Hint - Find Nm and then
estimate N and m separately). (25 marks)
(ii) Comment on your results and on the recovery so far – what do they indicate about the
reservoir drive? Provide a clear and reasoned discussion about what the material balance
equation has told you about the reservoir. (12 marks)
(iii) Discuss the reservoir management of this field. What problems are you likely to encounter
as the pressure falls further? What other recovery strategies might you consider? What
other information about the field would you need to know before making a final decision on
reservoir management? (13 marks)
k ro 0.8S o 0.30
2
o = 0.002 Pa.s
w = 0.001 Pa.s
= 0.15
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (20 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) The injection and production wells are 200 m apart. The average cross-sectional area of the
reservoir is 2,000 m2. Water is injected at a rate of 200 m3/day (measured at surface
conditions). At the reservoir pressure Bo = 1.5 and Bw = 0.98. Re-plot the graph in part (ii) as
oil produced (measured as surface volume in m3) against time (in days). (12 marks)
(iv) Estimate the time at which the oil production rate falls below 50 m3/day (measured at
surface conditions). (8 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2003
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
Pressure Gp Bg
(MPa) (106 scf) (rb/scf)
25.0 0 0.00167
24.5 45.8 0.00170
24.0 94.8 0.00174
23.5 143.4 0.00180
23.0 190.4 0.00186
22.5 234.0 0.00192
The initial water saturation is 0.25.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, estimate the original gas in place and the value of the aquifer
size times the aquifer compressibility. You may neglect the compressibility of the connate water
and rock in the gas field itself. (30 marks)
(ii) The residual gas saturation is 0.3. What is the water influx at the current reservoir pressure of
22.5 MPa? What fraction of the reservoir volume has been swept by water? What do expect to
happen when the pressure is dropped further? (20 marks)
(i) Is the reservoir normally pressured, over pressured or under pressured? (5 marks)
(ii) Find the depths of the oil/water and gas/oil contacts. Hence find the depth of the oil column.
(30 marks)
(iii) The areal extent of the reservoir is 1.6x10 6 m2 and the average porosity is 0.16. The oil
formation volume factor is 1.7. Find the oil volume in the reservoir, measured at surface
conditions if Swc = 0.2. (15 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
4. A core sample has the following relative permeabilities. (50 marks total)
k rw 0.25S w 0.2
2
k ro 0.9S o 0.3
2
o = 0.003 Pa.s
w = 0.001 Pa.s
= 0.25
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (20 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) The core has a cross-sectional area of 5 cm2 and a length of 15 cm. In a single-phase water
flow test the flow rate was 1 cm3/s for an imposed pressure drop across the core of 0.1 Mpa.
What is the permeability of the core? (8 marks)
(iv) With the same fixed injection rate of 1cm3/s use the results of part (ii) to find the volume of
oil recovered after 20 s of injection. (12 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2004
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
Pressure Gp Bg
(MPa) (108 scf) (rb/scf)
30.0 0 0.00201
29.5 970 0.00206
29.0 2066 0.00213
28.5 3244 0.00222
28.0 4530 0.00234
The initial water saturation is 0.3.
The material banace equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, estimate the original gas in place and the value of the
aquifer size times the aquifer compressibility. You may neglect the compressibility of the
connate water and rock in the gas field itself. (20 marks)
(iii) The residual gas saturation is 0.25. What will be the pressure at which water has swept
the entire reservoir? What will be the recovery factor? Hint – to do this you will need to
estimate one quantity – find a sensible estimate of this from an extrapolation of the data
given above. (13 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
k ro 0.8S o 0.3
2
o = 0.002 Pa.s
w = 0.0005 Pa.s
= 0.18
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (15 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) Explain the difference between pore volumes recovered and recovery factor. (5 marks)
(iv) You use these relative permeabilities to estimate recovery at the field scale. The initial oil in
place in the reservoir is 350 MMstb. You plan to inject a total of 50,000 stb water per day.
Plot oil recovery in stb against time in days. Bw = 1.1 and Bo = 1.55. (10 marks)
(v) What is the recovery factor after 10,000 days? (4 marks)
(vi) You perform a three-dimensional reservoir simulation of waterflooding and predict a
recovery factor of 0.3 after 10,000 days. Comment on how your answer compares with part
(v) above. (6 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2005
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
Pressure Gp Bg
(MPa) (108 scf) (rb/scf)
35.0 0 0.00345
34.5 87.4 0.00355
34.0 178 0.00367
33.5 266 0.00380
33.0 402 0.00411
The initial water saturation is 0.4.
Assuming a simple aquifer model, estimate the original gas in place and the value of the
aquifer size times the aquifer compressibility. You may neglect the compressibility of the
connate water and rock in the gas field itself. (17 marks)
The material balance equation is:
Bgi We
G p G1
B B
g g
(iv) The residual gas saturation is 0.3. What will be the pressure at which water has swept
the entire reservoir? What will be the recovery factor? To do this you will need to
estimate one quantity – find a sensible estimate of this based on the data given above.
(8 marks)
(v) What will happen once water has swept the entire reservoir? (5 marks)
2. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
k ro 0.8S o 0.3
2
o = 0.004 Pa.s
w = 0.001 Pa.s
= 0.12
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (15 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) The field has an estimated STOIIP of 400 MMstb. What is the pore volume of the reservoir?
Bo = 1.4. (7 marks)
(iv) Water is injected into several wells at a total rate of 300,000 stb/day. Plot oil recovery in stb
against time in days. Bw = 1.05. (10 marks)
(v) What is the recovery at 1,000 days? What is the recovery factor at 1,000 days? What is the
number of pore volumes produced? Why are the recovery factor and number of pore
volumes produced different? (8 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2006
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
Pressure Gp Bg
(MPa) (1012 scf) (rb/scf)
35 0 0.00245
34 0.079 0.00255
33 0.158 0.00270
32 0.228 0.00285
31 0.302 0.00316
The initial water saturation is 0.4.
Assuming a simple aquifer model, estimate the original gas in place and the value of the
aquifer size times the aquifer compressibility. You may neglect the compressibility of the
connate water and rock in the gas field itself. (17 marks)
(iv) The residual gas saturation is 0.25. What will be the pressure at which water has swept
the entire reservoir? What will be the recovery factor? To do this you will need to
estimate one quantity – find a sensible estimate of this based on the data given above.
(8 marks)
(v) Water has just broken through in the reservoir. What future problems are anticipated as
the pressure is dropped further? (5 marks)
(i) Draw schematic pressure-temperature phase diagrams for an oil and its associated gas in the
gas cap. Mark the temperature and pressure conditions on the diagram – why do they have
to be located where they are? (10 marks)
(ii) Is the reservoir normally pressured, over-pressured or under-pressured? (4 marks)
(iii) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column. (26
marks)
(iv) The areal extent of the reservoir is 50x106 m2 and the average porosity is 0.13 and the net-
to-gross ratio is 0.8. The oil formation volume factor is 1.41 rm3/sm3. The initial water
saturation is 0.3. Find the oil volume in the reservoir, measured at surface conditions. (10
marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(106 stb) (scf/stb) (scf/stb) (rb/stb) (MPa) (rb/scf)
0 - 700 1.612 40 0.000224
33.6 800 600 1.601 39 0.000289
55.5 1,800 475 1.587 38 0.000345
67.3 3,000 300 1.541 37 0.000407
(i) Discuss the types of process for which the material balance equation provides valuable
information and the cases where material balance is unlikely to be useful. (5 marks)
(ii) Is the field being produced above or below the bubble point? Explain your reasoning. (4
marks)
(iii) The material balance equation is:
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
What is the principal reservoir drive mechanism? There has been no water production and
you do not consider it likely that there is a strong aquifer drive. Estimate the initial oil in
place and the size of the gas cap, if any. You may assume that the compressibility of the
formation is negligible. (25 marks)
(iv) What is the recovery factor so far? Is this a good, bad or average recovery factor for this
process? (6 marks)
(v) What options would you consider for the further development of this field? What problems
are you likely to encounter? What extra information would you need? (10
marks)
k ro 0.8S o 0.25
2
o = 0.0032 Pa.s
w = 0.0008 Pa.s
= 0.12
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (17 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) The field has an estimated STOIIP of 200 MMstb. What is the pore volume of the
reservoir? Bo = 1.35. (5 marks)
(iv) Water is injected into several wells at a total rate of 50,000 stb/day. Plot oil recovery in stb
against time in days. Bw = 1.02. (10 marks)
(v) How much oil has been recovered at 10,000 days? What is the recovery factor at 10,000
days? Is this likely to be an over or under-estimate of the real recovery at this time? Explain
your answer. (8 marks)
UNIVERSITY OF LONDON
MSc. EXAMINATION 2007
For internal students of the Imperial College of Science, Technology and Medicine
Taken by students of the Department of Earth Science and Engineering
This paper is also taken for the relevant examination for the Associateship of the Royal School of Mines
Pressure Gp Bg
(MPa) (1010 scf) (rb/scf)
28 0 0.00578
27 1.06 0.00621
26 2.00 0.00667
25 2.88 0.00725
The initial water saturation is 0.3.
Assuming a simple aquifer model, estimate the original gas in place and the value of the
aquifer size times the aquifer compressibility. You may neglect the compressibility of the
connate water and rock in the gas field itself. (20 marks)
(v) The residual gas saturation is 0.3. What will be the pressure at which water has swept
the entire reservoir? What will be the recovery factor? To do this you will need to
estimate one quantity – find a sensible estimate of this based on the data given above.
(8 marks)
(vi) Water has just broken through in the reservoir. What future problems are anticipated as
the pressure is dropped further? (4 marks)
(i) What is drilling mud and what is it used for? What precautions need to be taken when drilling
a well into a hydrocarbon-bearing formation? (6 marks)
(ii) Is the reservoir normally pressured, over-pressured or under-pressured? (4 marks)
(iii) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column. (23
marks)
(iv) The areal extent of the reservoir is 1.7 km by 3.6 km. The average porosity is 0.15 and the
net-to-gross ratio is 0.76. The oil formation volume factor is 1.37 rm3/sm3. The initial water
saturation is 0.35. Find the oil volume in the reservoir, measured at surface conditions. (4
marks)
(v) Log analysis finds an oil/water contact that is 5 m above the contact estimated in part (iii).
Why is this? Which estimate would give the better estimate of initial oil volume? Use this
information to make an order-of-magnitude estimate of the reservoir permeability. Explain
your working. (13 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
k ro 0.8S o 0.3
2
o = 0.002 Pa.s
w = 0.001 Pa.s
= 0.15
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (17 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) Explain the difference between recovery factor and pore volumes recovered. (2 marks)
(iv) The field has an estimated STOIIP of 300 MMstb. What is the pore volume of the reservoir?
Bo = 1.45. (3 marks)
(v) Water is injected into several wells at a total rate of 300,000 stb/day. What real time (in
days) corresponds to one pore volume injected? Plot oil recovery in stb against time in days.
Bw = 1.04. (10 marks)
(vi) How much oil has been recovered at 2,000 days? What is the recovery factor at 5,000 days?
Is this likely to be an over or under-estimate of the real recovery at this time? Explain your
answer. (8 marks)
Pressure Gp Bg
(MPa) (106 scf) (rb/scf)
25.0 0 0.00167
24.0 45.8 0.00170
23.0 94.8 0.00174
22.0 143.4 0.00180
21.0 190.4 0.00186
20.0 234.0 0.00192
The initial water saturation is 0.25.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, estimate the original gas in place and the value of the aquifer
size times the aquifer compressibility. You may neglect the compressibility of the connate water
and rock in the gas field itself. (24 marks)
(iii) The residual gas saturation is 0.3. What is the water influx at the current reservoir pressure of 20
MPa? What fraction of the reservoir volume has been swept by water? What do expect to happen
when the pressure is dropped further? (16 marks)
(ii) What is drilling mud and what functions does it perform? (6 marks)
(iii) Is the reservoir normally pressured, over pressured or under pressured? (4 marks)
(iv) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column.
(12 marks)
(v) There is an error in the measurement of the gas pressure and the gauge reads 39.1 MPa.
What is your estimate of the depth of oil column now? Comment on your result.
(10 marks)
(vi) The areal extent of the reservoir is 3.5×10 6 m2 and the average porosity is 0.15. The oil
formation volume factor is 1.7. The average water saturation in the oil zone is 0.2. Find the
oil volume in the reservoir (using the original estimate of the depth of the oil column),
measured at surface conditions. (6 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(107 stb) (scf/stb) (scf/stb) (rb/stb) (MPa) (rb/scf)
k ro 0.8So 0.25
4
o = 0.002 Pa.s
w = 0.001 Pa.s
= 0.15
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (15 marks)
(ii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iii) Explain the difference between pore volumes recovered and recovery factor. (5 marks)
(iv) You use these relative permeabilities to estimate recovery at the field scale. The initial oil in
place in the reservoir is 450 MMstb. You plan to inject a total of 90,000 stb water per day.
Plot oil recovery in stb against time in days. Bw = 1.05 and Bo = 1.45. (10 marks)
(v) What is the recovery factor after 4,000 days? (4 marks)
(vi) You perform a three-dimensional reservoir simulation of waterflooding and predict a
recovery factor of 0.3 after 4,000 days. Comment on your answer compared with part (v).
(6 marks)
Pressure Gp Bg
(MPa) (106 scf) (rb/scf)
32.00 0 0.00089
31.75 358 0.00095
31.50 684 0.00102
31.25 968 0.00110
The initial water saturation is 0.31.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, estimate the original gas in place and the value of the aquifer
size times the aquifer compressibility. You may neglect the compressibility of the connate water
and rock in the gas field itself. (20 marks)
(iii) At a pressure of 31.25 MPa there is significant water production. Estimate the average gas
saturation in the field. Comment on this value. Is there likely to be significant future production?
Why is there significant water production? What further development options would you
consider in this field? (20 marks)
(i) Explain what precautions need to be made when drilling a well through a hydrocarbon-
bearing formation. Why is a drilling mud used? (4 marks)
(ii) Is the reservoir normally pressured, over pressured or under pressured? (4 marks)
(iii) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column. (24
marks)
(iv) The reservoir structure can be approximated as a spherical dome of radius 5,000 m with the
top of the field at a depth of 3,050 m. The average porosity is 0.21, the net-to-gross is 0.85
and the oil formation volume factor is 1.65. The average water saturation in the oil zone is
0.28. Estimate the oil volume in the reservoir, measured at surface conditions. (18 marks)
3. You have the following data for a large oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(106 stb) (scf/stb) (scf/stb) (rb/stb) (MPa) (rb/scf)
Pressure Gp Bg
(MPa) (106 scf) (rb/scf)
31.0 0 0.00123
30.6 25.0 0.00137
30.2 47.4 0.00159
29.8 66.4 0.00203
The initial water saturation is 0.28 and the residual gas saturation, from laboratory measurements,
is 0.32.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself.
Is there evidence of significant water influx? (20 marks)
(iii) What is the recovery factor now? Estimate the maximum possible recovery factor for this field
based on the laboratory measurement of residual gas saturation. Comment on your answer. What
further development options would you consider in this field? (20 marks)
(i) Explain physically why gas and oil pressures in a reservoir are typically higher than the water
pressure in the surrounding aquifer. (8 marks)
(ii) Is the reservoir normally pressured, over pressured or under pressured? (4 marks)
(iii) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column. (22
marks)
(iv) The horizontal cross-section of the reservoir through the oil zone is approximately an ellipse
with a maximum diameter of 3,450m and a minimum diameter of 1,680m. The average
porosity is 0.24, the net-to-gross is 0.79 and the oil formation volume factor is 1.43. The
average water saturation in the oil zone is 0.34. Estimate the oil volume in the reservoir,
measured at surface conditions. (16 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
(106 stb) (scf/stb) (scf/stb) (rb/stb) (MPa) (rb/scf)
k ro 0.9S o 0.1
4
o = 0.002 Pa.s
w = 0.0005 Pa.s
= 0.2
(i) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (12 marks)
(ii) Is this likely to be a water-wet, oil-wet or mixed-wet reservoir? Explain your answer carefully.
What definitive test could you use to determine wettability? (8 marks)
(iii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iv) You use these relative permeabilities to estimate recovery at the field scale. The initial oil in
place in the reservoir is 400 MMstb. You plan to inject a total of 10,000 stb water per day for
6,000 days. What is the volume of oil produced and the recovery factor? Bw = 1.01 and Bo =
1.25. (10 marks)
(v) Discuss how you would use these results in combination with a reservoir simulation study to
predict recovery and design an optimal injection scheme. How would you estimate the
sweep efficiency? (10 marks)
(ii) You have the following production data for a dry gas field that might be produced by a natural
water drive.
Pressure Gp Bg
(MPa) (106 scf) (rb/scf)
33.0 0 0.00089
32.6 290 0.00098
32.2 568 0.00113
31.8 814 0.00130
The initial water saturation is 0.26 and the residual gas saturation, from laboratory measurements,
is 0.27.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? Estimate the
maximum possible recovery factor for this field based on the laboratory measurement of residual
gas saturation. (12 marks)
(iv) It is suggested that this field could be used for gas storage and that CO2 or nitrogen could be
injected near the gas/water contact as a ‘cushion gas’. Comment on this proposal. What do you
think a ‘cushion gas’ is and what role does it play? (10 marks)
2. You are advised to draw a large, clear sketch to illustrate your work. You will be awarded
marks for this sketch even if your calculations are in error. (50 marks total)
(i) Write down the definition of isothermal compressibility. Write the definition in terms of
density. (4 marks)
(ii) An aquifer is used for CO2 storage. An injection well and a monitoring wells are drilled and
pressures are taken at different depths. The aquifer brine has a density of 1,120 kg.m-3.
Before any injection the following pressure measurements are taken. Find the depth of the
monitoring well measurement. (6 marks)
Well 1 - injection. Depth = 1,120 m. Pressure = 11.2 MPa.
Well 2 - monitoring. Depth = ??. Pressure = 12.1 MPa.
(iii) After the injection of 1 million tonnes (10 9 kg) of CO2 the following measurements are made.
What is the depth of the CO2-water contact? At reservoir conditions, the CO2 has a density
of 600 kg.m-3. Why has the pressure in the monitoring well increased? (20 marks)
Well 1 – injection of CO2. Depth = 1,120 m. Pressure = 13.1 MPa.
Well 2 – monitoring (brine). Depth = answer to part (ii). Pressure = 13.7 MPa.
(iv) CO2 has a constant compressibility of 10 -8 Pa-1. Using the definition of compressibility, write
an expression for density as a function of pressure. Derive an expression for pressure as a
function of depth. Make a rough estimate of the error in depth associated with the
assumption of incompressibility. (20 marks)
3. You have the following data for an oil reservoir. (50 marks total)
Np Rp Rs Bo P Bg
k ro 0.6S o 0.25
2
o = 0.0015 Pa.s
w = 0.0005 Pa.s
= 0.25
(i) Define relative permeability from the multiphase Darcy equation. Discuss how relative
permeability affects overall field-scale oil production. What features of the relative
permeability curves have a key impact on recovery? (18 marks)
(ii) Plot the fractional flow curve and from this draw a plot of water saturation against
dimensionless velocity for this system. Indicate clearly the shock front saturation and
dimensionless speed. (12 marks)
(iii) Plot a graph of pore volumes recovered against pore volumes injected. (10 marks)
(iv) You use these relative permeabilities to estimate recovery at the field scale. The initial oil in
place in the reservoir is 200 MMstb. You plan to inject a total of 45,000 stb water per day for
7,000 days. What is the volume of oil produced and the recovery factor? Bw = 1.02 and Bo =
1.30. (10 marks)
Gp Pressure Bg
(106 scf) (MPa) (rb/scf)
0 34 0.00234
72 33 0.00298
112 32 0.00345
129 31 0.00354
155 30 0.00399
The initial water saturation is 0.23 and the residual gas saturation, from laboratory measurements,
is 0.41.
The material balance equation is:
Bgi We
G p G1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? Estimate the
maximum possible recovery factor for this field based on the laboratory measurement of residual
gas saturation. Comment on your values. (12 marks)
(iv) Does this field have a strong water drive? Discuss the advantages and disadvantages of a natural
water drive in a gas field. (10 marks)
1
(d)
0.8
kr 0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
sw
(ii) You have the following production data for a dry gas field that might be produced by a natural
water drive.
Gp Pressure Bg
(106 scf) (MPa) (rb/scf)
0 32 0.000134
26 31 0.000146
63 30 0.000176
85 29 0.000201
105 28 0.000228
The initial water saturation is 0.29 and the residual gas saturation, from laboratory measurements,
is 0.31.
The material balance equation is:
Bgi We
G p G1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? What is the
pressure at which water influx will have invaded the whole field? Estimate the maximum possible
recovery factor for this field based on the laboratory measurement of residual gas saturation.
(12 marks)
(iv) It is suggested that this gas field is used to store carbon dioxide. Comment on this suggestion.
How could the injected carbon dioxide help or hinder production rates and ultimate
recovery? (10 marks)
2. You measure the following from appraisal wells in an oilfield: (50 marks total)
Depth (m) Pressure (MPa) Fluid and density (kg.m-3)
1,235 12.32 Gas, 250
1,356 12.67 Oil, 750
1,467 13.79 Water, 1,055
The acceleration due to gravity = 9.81 ms-2. Depths are measured from the surface.
(i) Explain physically why gas and oil pressures in a reservoir are typically higher than the water
pressure in the surrounding aquifer. (8 marks)
(ii) Is the reservoir normally pressured, over pressured or under pressured? Explain your answer
carefully. (6 marks)
(iii) Find the depths of the oil/water and gas/oil contacts and the depth of the oil column. (18
marks)
(iv) Define what is meant by primary and secondary production. (4 marks)
(v) Later, after a brief period of primary production the pressures are 11.54 MPa, 11.79 MPa and
13.72 MPa for the gas, oil and water respectively. Explain carefully what this indicates
concerning the mechanism for oil production in the field. What secondary production
mechanisms would you recommend? (14 marks)
4. Material balance for an oilfield. You are given the following data for an oilfield. (50 marks
total)
Np Gp Rs Bo Bg
(MMstb) (MMscf) P (MPa) (scf/stb) (rb/stb) (rb/scf)
0 0 32 400 1.356 0.000187
1.41 480 30 400 1.361 0.000199
1.98 1568 28 370 1.355 0.000213
3.41 3016 26 345 1.349 0.000251
5.78 6890 24 295 1.335 0.000302
(i) Define, physically, the process of gas gravity drainage. During what recovery processes does
it occur? Why can it provide a very high local displacement efficiency? (10 marks)
(ii) From the material balance equation below, find the size of the oilfield and the relative size
of the gas cap (if any). You may assume that there is no active aquifer and can ignore the
compressibility for the formation. (18 marks)
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
(iii) From comparing the expansion of oil to the expansion of gas (for the final set of data),
quantify relatively much recovery is contributed from gas expansion and how much from oil
expansion. Comment on your answer. (12 marks)
(iv) What is the recovery factor now? Is this good as an ultimate recovery factor for this field?
Comment on the relative pressure decline in the field. What further development options
would you consider? (10 marks)
Gp Pressure Bg
(106 scf) (MPa) (rb/scf)
0 41 0.000203
60 40 0.000224
118 39 0.000255
170 38 0.000305
208 37 0.000362
The initial water saturation is 0.24 and the residual gas saturation, from laboratory measurements,
is 0.29.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? What is the
pressure at which water influx will have invaded the whole field? Estimate the maximum possible
recovery factor for this field based on the laboratory measurement of residual gas saturation.
(12 marks)
(iv) The field also produces a small amount of condensate (oil). When Gp=208×106 scf, Np, the oil
produced is 50,000 stb. What does this indicate about the field? How might this impact the
management of the field? (10 marks)
2. You measure the following from appraisal wells in an oilfield: (50 marks total)
Depth (m) Pressure (MPa) Fluid and density (kg.m-3)
2,250 17.51 Gas, 305
2,285 17.63 Oil, 650
2,327 18.01 Water, 1,040
The acceleration due to gravity = 9.81 ms-2. Depths are measured from the surface.
(i) Explain physically why gas and oil pressures in a reservoir are typically higher than the water
pressure in the surrounding aquifer. (8 marks)
(ii) Is the reservoir normally pressured, over pressured or under pressured? Explain your
answer. (4 marks)
(iii) Find the depths of the free oil and free water levels. Hence find the height of the oil column.
(18 marks)
(iv) Log measurements place the oil/water contact at a depth of 2,290 m. Explain the
discrepancy with the answer in part (iii). Which value would you take to calculate the original
oil in place? What is your estimate now of the height of the oil column? (12 marks)
(v) Find the original oil in place, if the reservoir has an area of 3.6×10 7 m2, an average porosity
of 0.22, a net-to-gross of 0.85 and an oil formation volume factor of 1.41. You may quote
the answer in sm3. The initial water saturation is 0.25. (8 marks)
krw
Sw 0.24
0.73
kro
So 0.3
3
0.72
o = 0.002 Pa.s
w = 0.001 Pa.s
(iii) Calculate the saturation as a function of dimensionless velocity and the pore volumes
produced as a function of pore volumes injected. Plot your answers on a graph. (15 marks)
(iv) Plot the oil produced as a function of time if Bo=1.5, Bw=1.1, the total water injection rate is
105 stb/day. The pore volume of the reservoir is 1.5×10 8 rb. (10 marks)
4. Material balance for an oilfield. You are given the following data for an oilfield. (50 marks total)
Np Gp Rs Bo
(MMstb) (MMscf) P (MPa) (scf/stb) (rb/stb) Bg (rb/scf)
0 0 32 400 1.356 0.000265
1.67 480 30 400 1.361 0.000321
2.67 1230 28 370 1.355 0.000355
3.01 1900 26 345 1.349 0.000365
4.21 2600 24 295 1.335 0.000398
(i) Explain physically why – in general – you inject water into the aquifer (or the base of the oil
column) and gas into the gas cap (or the top of the field) in secondary production. (10
marks)
(ii) From the material balance equation below, find the size of the oilfield and the relative size
of the gas cap (if any). You may assume that there is no active aquifer and can ignore the
compressibility for the formation. (18 marks)
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
(iii) From comparing the expansion of oil to the expansion of gas (for the final set of data),
quantify relatively how much recovery is contributed by gas expansion and how much from
oil expansion. Comment on your answer. (12 marks)
(iv) What is the recovery factor now? Is this as good as a typical ultimate recovery factor for this
type of field? Comment on the relative pressure decline in the field. What further
development options would you consider? (10 marks)
Gp Pressure Bg
(106 scf) (MPa) (rb/scf)
0 45 0.0035
30 44 0.004
56 43 0.0047
80 42 0.0058
95 41 0.0072
40 0.0088
39 0.0101
The initial water saturation is 0.21 and the residual gas saturation, from laboratory measurements,
is 0.32.
The material balance equation is:
Bgi We
G p G 1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? What is the
pressure at which water influx will have invaded the whole field? Estimate the maximum possible
recovery factor for this field based on the laboratory measurement of residual gas saturation.
Comment on your answer. (12 marks)
2. You measure the following from appraisal wells in an oilfield: (50 marks total)
(i) Is the reservoir normally pressured, over pressured or under pressured? Explain your
answer. (4 marks)
(iii) Find the depths of the free oil and free water levels. Hence find the height of the oil column.
(18 marks)
(iii) The field is produced by water injection into the aquifer. After 5 years of production the
following measurements are made.
Depth (m) Pressure (MPa) Fluid and density (kg.m-3)
3,500 32.12 Gas, 420
3,720 33.08 Oil, 520
3,857 38.85 Water, 1,030
What do these measurements imply about the connectivity of the field? What concerns
would you have associated with continued water injection? (18 marks)
(iv) Discuss options for further development of the field. Take note of the gas and oil densities.
What type of field is this likely to be? (10 marks)
k rw
S w 0.33
0.16
k ro
S o 0 .32
0 .4
o = 0.003 Pa.s
w = 0.001 Pa.s
(iv) Calculate the saturation as a function of dimensionless velocity and the pore volumes
produced as a function of pore volumes injected. Plot your answers on a graph. (15 marks)
(v) Plot the oil produced as a function of time if Bo=1.4, Bw=1.05, and the total water injection
rate is 2×105 stb/day. The pore volume of the reservoir is 3.5×10 8 rb. (10 marks)
4. Material balance for an oilfield. You are given the following data for an oilfield that is being produced
under primary production. (50 marks total)
Np Gp Rs Bo Bg
(MMstb) (MMscf) P (MPa) (scf/stb) (rb/stb) (rb/scf)
0 0 25 700 1.421 0.000367
1.48 1230 24 700 1.432 0.000426
2.45 2460 23 600 1.397 0.000456
5.75 5800 22 450 1.361 0.000521
8.62 10980 21 312 1.342 0.000555
(i) Define carefully primary, secondary and tertiary production. What is the difference between
EOR and tertiary production? (10 marks)
(ii) From the material balance equation below, find the size of the oilfield and the relative size
of the gas cap (if any). You may assume that there is no active aquifer and can ignore the
compressibility for the formation. (18 marks)
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
(iii) From comparing the expansion of oil to the expansion of gas (for the final set of data),
quantify relatively how much recovery is contributed by gas expansion and how much from
oil expansion. (7 marks)
(iv) Comment on the values of Rp: can Rp be less than
Rs?
(5 marks)
(v) What is the recovery factor now? Is this as good as a typical ultimate recovery factor for this
type of field? Comment on the relative pressure decline in the field. What further
development options would you consider? (10 marks)
2. You measure the following from appraisal wells in an oilfield: (50 marks total)
Depth (m) Pressure (MPa) Fluid and density (kg.m-3)
2,500 34.51 Gas, 360
2,620 35.08 Oil, 634
2,857 37.21 Water, 1,067
The acceleration due to gravity = 9.81 ms-2. Depths are measured from the surface.
(i) Is the reservoir normally pressured, over pressured or under pressured? Explain your
answer. (4 marks)
(ii) Find the depths of the free oil and free water levels. Hence find the height of the oil column.
(18 marks)
(iii) Log measurements in the field suggest the presence of water between 2,590 m and 2,610 m,
and below 2,695 m. How can you reconcile this with the results in part (ii)? (14 marks)
(iv) The average cross-sectional area of the field is 10 km2, while the average porosity and water
saturation in the hydrocarbon-bearing zones is 0.21 and 0.32 respectively. The oil formation
volume factor is 1.5. Estimate the original oil in place and explain your working clearly. (14
marks)
o = 0.002 Pa.s
w = 0.001 Pa.s
(iv) Calculate the saturation as a function of dimensionless velocity and the pore volumes
produced as a function of pore volumes injected. Plot your answers on a graph. (15 marks)
4. Material balance for an oilfield. You are given the following data for an oilfield that is being produced
under primary production. (50 marks total)
Np Gp Rs Bo
(MMstb) (MMscf) P (MPa) (scf/stb) (rb/stb) Bg (rb/scf)
0 0 32 400 1.356 0.000187
1.41 480 30 400 1.361 0.000199
1.98 1568 28 370 1.355 0.000213
3.41 3016 26 345 1.349 0.000251
5.78 6890 24 295 1.335 0.000302
(i) In this field, are we producing above or below the bubble point and why? (4 marks)
(ii) Define R, Rs and Rp. (6 marks)
(iii) From the material balance equation below, find the size of the oilfield and the relative size
of the gas cap (if any). You may assume that there is no active aquifer and can ignore the
compressibility for the formation. (18 marks)
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
(iv) Compare the expansion of oil to the expansion of gas (for the final set of data), to quantify
relatively how much recovery is contributed by gas expansion and how much from oil
expansion. (7 marks)
(v) What options are there for dealing with the produced gas? (5
marks)
(vi) What is the recovery factor now? Is this as good as a typical ultimate recovery factor for this
type of field? Comment on the relative pressure decline in the field. What further
development options would you consider? (10 marks)
Gp Pressure Bg
(106 scf) (MPa) (rb/scf)
0 40 0.000521
100 39 0.000545
220 38 0.000612
300 37 0.000639
370 36 0.000687
The initial water saturation is 0.15 and the residual gas saturation, from laboratory measurements,
is 0.26.
The material balance equation is:
Bgi We
G p G1
B B
g g
Assuming a simple aquifer model, make an approximate estimate of the original gas in place and
the value of the aquifer size times the aquifer compressibility. You may neglect the compressibility
of the connate water and rock in the gas field itself. (18 marks)
(iii) What is the recovery factor now? What is the average gas saturation in the field? What fraction
of the recovery to date can be attributed to water drive, and what to gas expansion? What is the
pressure drop necessary for water influx to invade the whole field? Comment on these answers
carefully: how should the field be managed? (20 marks)
2. You measure the following from appraisal wells in an oilfield: (50 marks total)
Depth (m) Pressure (MPa) Fluid and density (kg.m-3)
3,500 34.21 Gas, 420
3,720 35.21 Oil, 520
3,857 36.34 Water, 1,030
The acceleration due to gravity = 9.81 ms-2. Depths are measured from the
surface.
(i) Is the reservoir normally pressured, over pressured or under pressured? Explain your
answer. (4 marks)
(ii) Find the depths of the free oil and free water levels. Hence find the height of the oil column.
(18 marks)
(iii) Explain why the oil/water contact is often found above the free water level. In this example
the oil/water contact lies 2 m above the free water level. What is the true height of the oil
column? (8 marks)
(iv) Explain the concept of net-to-gross. How is this estimated? (8 marks)
(v) The average cross-sectional area of the field is 13 km2, while the average porosity and water
saturation in the hydrocarbon-bearing zones are 0.19 and 0.26 respectively. The oil
formation volume factor is 1.4. The net-to-gross is 0.8. Estimate the original oil in place.
Comment on the size of the field? How would this be classified? (12 marks)
4. Material balance for an oilfield. You are given the following data for an oilfield that is being produced
under primary production. (50 marks total)
Np Gp Rs Bo
(MMstb) (MMscf) P (MPa) (scf/stb) (rb/stb) Bg (rb/scf)
0 0 30 550 1.214 0.000218
16.01 850 29 500 1.207 0.000229
45.12 1568 28 430 1.198 0.000246
66.54 3016 27 400 1.195 0.000259
82.03 6890 26 380 1.193 0.00027
(i) From the material balance equation below, find the size of the oilfield and the relative size
of the gas cap (if any). Also quote the gas initially in place, including solution gas. You may
assume that there is no active aquifer and can ignore the compressibility for the
formation. (20 marks)
N p Bo R p Rs Bg
Bo Boi Rsi Rs Bg
Boi
NBoi
B c S cf
m g 1 (1 m) w wc P
Bgi 1 S
wc
We W p Bw
(ii) Compare the expansion of oil to the expansion of gas (for the final set of data), to quantify
relatively how much recovery is contributed by gas expansion and how much from oil
expansion. Comment on the result. (7 marks)
(iii) Before production started, the most-likely case estimate of the oil and gas in place were
1,000 MMstb and 5,000 MMMscf respectively. Comment on these values and how they
compare with the answers to part (i). How might the discrepancies be reconciled? (13 marks)
(iv) What is the recovery factor now? Is this as good as a typical ultimate recovery factor for this
type of field? Comment on the relative pressure decline in the field. What further
development options would you consider? Take into account the size of the field. (10 marks)
Use this information to plot a graph of water saturation versus height above the oil/water
contact in the reservoir (as in Part (ii), but with real numbers!). The water (brine density) =
1050 kgm-3, the oil density = 750 kgm-3, the acceleration due to gravity = 9.81 ms-2. The
porosity of the rock sample = 0.25 and the permeability is 500 mD. The average porosity of
the reservoir is 0.2 and the average permeability is 200 mD. (11 marks)
Flow
direction
30o
HYDROGEOLOGY EXAM
Answer any FIVE questions
(i) Draw a diagram showing the zones of subsurface water. Define the water table and the
capillary fringe. (8 marks)
(ii) A well is being drilled. At a depth of 6m below the ground surface moist sandy silt is
encountered. At 10m there us standing water in the hole. What do these results mean
in terms of the zones defined in part (i) above? (4 marks)
(iii) A pressure measurement is made at a depth of 12m. The measured pressure is 30 kPa
above atmospheric pressure. What can be said about the direction of groundwater flow?
(8 marks)
(i) You have an unconfined aquifer. Define the irreducible water saturation. Explain how,
by taking a sample of the aquifer sediment, the irreducible water saturation can be
measured. (6 marks)
(ii) The aquifer has an effective porosity of 0.25 and an irreducible water saturation of 0.2.
The aquifer depth is 40m and its area is 1.4km2. What is the water volume contained in
the aquifer if it is completely saturated? What is the specific yield of the aquifer? How
much water can be extracted from the aquifer if it were completely drained? (10 marks)
(iii) The annual rainfall is 600mm. Neglecting evaporation or run-off, how long would it take
to recharge the aquifer completely? (4 marks)
(i) Discuss the issues to be addressed when assessing water quality in an unconfined aquifer.
(10 marks)
(ii) What samples should be taken, how can you ensure that they are representative and
what significance would you attach to different measurements? (10 marks)
END OF EXAM
(i) Define the following terms and give appropriate units: specific yield, coefficient of storage
and transmissivity. (6 marks)
(ii) An unconfined aquifer has an area of 10,000 m2 and a depth of 15 m. How much water can
be released from the aquifer if the specific yield is 0.15. (10 marks)
(iii) The porosity is 0.4. What is the connate or irreducible water saturation? (4 marks)
8. Relative permeability and the effect of flow rate. (20 marks total)
(i) Write down the multiphase Darcy law and define relative permeability. (5 marks)
(ii) Define the capillary number. What does it represent physically? (5 marks)
(iii) Draw schematic figures that show the effect of capillary number on relative permeability and
residual oil saturation. Explain the figures. (10 marks)
END OF EXAM
(i) Define coefficient of storage. Explain the units used to measure it. (5 marks)
(ii) What happens physically when the pressure is dropped in a confined aquifer for which the
coefficient of storage is defined? Why is the coefficient of storage much lower than the
specific yield in an unconfined aquifer? (10 marks)
(iii) A confined and fully saturated aquifer has a coefficient of storage of 0.001. How much water
is released from an aquifer of area 2000m2 for a drop in head of 2m?
(5 marks)
END OF EXAM
(i) Write an equation that relates the capillary pressure to the Leverett J function. Define
all the terms and give appropriate units. (5 marks)
(ii) In the field the average permeability is 100 mD, the porosity is 0.2 and the interfacial
tension is 30 mN/m. Plot a graph of water saturation against height above the free water
level in the reservoir. The oil density is 800 kgm-3 and the brine density is 1100 kgm-3.
g=9.81 ms-2. (15 marks)
(iii) What approximations have you made in this analysis? (5 marks)
END OF EXAM
(i) Write an equation that relates the capillary pressure to the Leverett J function. Define all
the terms and give appropriate units. (5 marks)
(ii) In the field the average permeability is 200 mD, the porosity is 0.15 and the interfacial tension
is 25 mN/m. Plot a graph of water saturation against height above the free water level in the
reservoir. The oil density is 700 kgm-3 and the brine density is 1050 kgm-3. g=9.81 ms-2. (15
marks)
(iii) What approximations have you made in this analysis and what did you have to assume? (5
marks)
Oil
2r Water
END OF EXAM
(i) Write an equation that relates the capillary pressure to the Leverett J function. Define all the
terms and give appropriate units. (5 marks)
(ii) In the field the average permeability is 1 mD, the porosity is 0.15 and the interfacial tension
is 25 mN/m. Plot a graph of capillary pressure as a function of water saturation for the field.
(10 marks)
(iii) The reservoir is waterflooded. The matrix is surrounded by fractures, forming blocks
approximately 6 m tall. The brine density is 1,150 kgm-3 and the oil density is 800 kgm-3.
Estimate the final water (brine) saturation in the matrix. (10 marks)
5. Conservation equations for CO2 storage in a fractured medium. (25 marks total)
Huge fractured aquifers are possible storage locations for CO2 collected from power stations and
other industrial plants. The CO2 flows through the fractures. CO2 also dissolves in brine – this CO2
saturated brine can enter the matrix.
(i) The CO2 in its own phase remains in the fractures. Explain physically what prevents the CO 2
in its own phase entering the matrix. (5 marks)
(ii) By what physical mechanism does the CO2 dissolved in brine move through the matrix? (4
marks)
(iii) If the fractures are closely spaced, then all the water in the matrix in contact with a fracture
containing CO2 will have the same dissolved CO2 concentration, equal to the solubility. If this
solubility is Cs, write down a conservation equation for the flow of CO2. You may assume that
the only Darcy flow is in the fractures and can ignore water in the fractures themselves. To
simplify the analysis, you can assume the Darcy flow of CO2 is Scqt, where Sc is the saturation
and qt is the total (Darcy) velocity. f is the porosity of the fractures and m is the porosity of
the matrix. Derive an expression for the speed of the CO2 if the fracture saturation is 1. Draw
a sketch to illustrate how the CO2 moves that explains your answer. (13 marks)
(iv) Find the speed of the CO2 in the fractures with dissolution if the total, Darcy velocity is 10-7
ms-1, f = 0.05%, m = 30%. Cs = 40 kgm-3 and the density of CO2 in its own phase is 600 kg.m-
3
. (3 marks)
END OF EXAM
(v) In this field, water is injected to displace oil. If the oil and water viscosities are similar,
estimate – approximately – the water saturation at which more water than oil will be
produced from the field. What fraction of the oil that was originally in the reservoir will be
produced? (7 marks)
(i) Write an equation that relates the capillary pressure to the Leverett J function. Define all
the terms and give appropriate units. (5 marks)
(ii) In the field the average permeability is 20 mD, the porosity is 0.15 and the interfacial tension
is 20 mN/m. Plot a graph of capillary pressure as a function of water saturation for the field.
(10 marks)
(iii) Is the core sample water-wet, oil-wet or mixed-wet? Explain your answer. What is the the
Amott wettability index for water? (10 marks)
END OF EXAM