Sei sulla pagina 1di 487

Introduction and Case Studies

CONTENTS
1

WHAT IS A SIMULATION MODEL?


1.1 A Simple Example of a Simulation Model
1.2 A Note on Units

WHAT IS A RESERVOIR SIMULATION


MODEL?
2.1 The Task of Reservoir Simulation
2.2 What Are We Trying To Do and How Complex
Must Our Model Be?

FIELD APPLICATIONS OF RESERVOIR


SIMULATION
3.1 Reservoir Simulation at Appraisal and in
Mature Fields
3.2 Introduction to the Field Cases
3.3 Case 1: The West Seminole Field Simulation
Study (SPE10022, 1982)
3.4 Ten Years Later - 1992
3.5 Case 2: The Anguille Marine Simulation
Study (SPE25006, 1992)
3.6 Case 3: Ubit Field Rejuvenation
(SPE49165,1998)
3.7 Discussion of Changes in Reservoir
Simulation; 1970s - 2000
3.8 The Treatment of Uncertainty in Reservoir
Simulation

STUDY EXAMPLE OF A RESERVOIR


SIMULATION

TYPES OF RESERVOIR SIMULATION


MODEL
5.1 The Black Oil Model
5.2 More Complex Reservoir Simulation
Models
5.3 Comparison of Field Experience with
Various Simulation Models

SOME FURTHER READING ON RESERVOIR


SIMULATION

APPENDIX A - References
APPENDIX B - Some Overview Articles on
Reservoir Simulation
1. Reservoir Simulation: is it worth the effort?
SPE Review, London Section monthly panel
discussion November 1990.
2. The Future of Reservoir Simulation - C.
Galas, J. Canadian Petroleum Technology, 36,
January 1997.
3. What you should know about evaluating
simulation results - M. Carlson; J. Canadian
Petroleum Technology, Part I - pp. 21-25,
36, No. 5, May 1997; Part II - pp. 52-57, 36,
No. 7, August 1997.

LEARNING OBJECTIVES:
Having worked through this chapter the student should:

Be able to describe what is meant by a simulation model, saying what analytical


models and numerical models are.

Be familiar with what specifically a reservoir simulation model is.

Be able to describe the simplifications and issues that arise in going from the
description of a real reservoir to a reservoir simulation model.

Be able to describe why and in what circumstances simple or complex reservoir


models are required to model reservoir processes.

Be able to list what input data is required and where this may be found.

Be able to describe several examples of typical outputs of reservoir simulations


and say how these are of use in reservoir development.

Know the meaning of all the highlighted terms - or terms referred to in the
Glossary - in Chapter 1 e.g. history matching, black oil model, transmissibility,
pseudo relative permeability etc.

Be able to describe and discuss the main changes in reservoir simulation over
the last 40 years from the 60's to the present - and say why these have
occurred.

Know in detail and be able to compare the differences between what


reservoir simulations can do at the appraisal and in the mature stages of reservoir
development.

Have an elementary knowledge of how uncertainty is handled in reservoir


simulation.

Know all the types of reservoir simulation models and what type of problem
or reservoir process each is used to model.

Know or be able to work out the equations for the mass of a phase or component
in a grid block for a black oil or compositional model.

Introduction and Case Studies

CONTENTS
1

WHAT IS A SIMULATION MODEL?


1.1 A Simple Example of a Simulation Model
1.2 A Note on Units

WHAT IS A RESERVOIR SIMULATION


MODEL?
2.1 The Task of Reservoir Simulation
2.2 What Are We Trying To Do and How Complex
Must Our Model Be?

FIELD APPLICATIONS OF RESERVOIR


SIMULATION
3.1 Reservoir Simulation at Appraisal and in
Mature Fields
3.2 Introduction to the Field Cases
3.3 Case 1: The West Seminole Field Simulation
Study (SPE10022, 1982)
3.4 Ten Years Later - 1992
3.5 Case 2: The Anguille Marine Simulation
Study (SPE25006, 1992)
3.6 Case 3: Ubit Field Rejuvenation
(SPE49165,1998)
3.7 Discussion of Changes in Reservoir
Simulation; 1970s - 2000
3.8 The Treatment of Uncertainty in Reservoir
Simulation

STUDY EXAMPLE OF A RESERVOIR


SIMULATION

TYPES OF RESERVOIR SIMULATION


MODEL
5.1 The Black Oil Model
5.2 More Complex Reservoir Simulation
Models
5.3 Comparison of Field Experience with
Various Simulation Models

SOME FURTHER READING ON RESERVOIR


SIMULATION

APPENDIX A - References
APPENDIX B - Some Overview Articles on
Reservoir Simulation
1. Reservoir Simulation: is it worth the effort?
SPE Review, London Section monthly panel
discussion November 1990.
2. The Future of Reservoir Simulation - C.
Galas, J. Canadian Petroleum Technology, 36,
January 1997.
3. What you should know about evaluating
simulation results - M. Carlson; J. Canadian
Petroleum Technology, Part I - pp. 21-25,
36, No. 5, May 1997; Part II - pp. 52-57, 36,
No. 7, August 1997.

LEARNING OBJECTIVES:
Having worked through this chapter the student should:

Be able to describe what is meant by a simulation model, saying what analytical


models and numerical models are.

Be familiar with what specifically a reservoir simulation model is.

Be able to describe the simplifications and issues that arise in going from the
description of a real reservoir to a reservoir simulation model.

Be able to describe why and in what circumstances simple or complex reservoir


models are required to model reservoir processes.

Be able to list what input data is required and where this may be found.

Be able to describe several examples of typical outputs of reservoir simulations


and say how these are of use in reservoir development.

Know the meaning of all the highlighted terms - or terms referred to in the
Glossary - in Chapter 1 e.g. history matching, black oil model, transmissibility,
pseudo relative permeability etc.

Be able to describe and discuss the main changes in reservoir simulation over
the last 40 years from the 60's to the present - and say why these have
occurred.

Know in detail and be able to compare the differences between what


reservoir simulations can do at the appraisal and in the mature stages of reservoir
development.

Have an elementary knowledge of how uncertainty is handled in reservoir


simulation.

Know all the types of reservoir simulation models and what type of problem
or reservoir process each is used to model.

Know or be able to work out the equations for the mass of a phase or component
in a grid block for a black oil or compositional model.

Introduction and Case Studies

BRIEF DESCRIPTION OF CHAPTER 1


A brief overview of Reservoir Simulation is first presented. This module then
develops this introduction by going straight into three field examples of applied
simulation studies. This is done since this course has some reservoir engineering
pre-requisites which will have made the student aware of many of the issues in
reservoir development. In these literature examples, we introduce many of the
basic concepts that are developed in detail throughout the course e.g. gridding
of the reservoir, data requirements for simulation, well controls, typical outputs
from reservoir simulation (cumulative oil, watercuts etc.), history matching and
forward prediction etc. After briefly discussing the issue of uncertainty in
reservoir management, some calculated examples are given. Finally, the
various types of reservoir simulation model which are available for calculating
different types of reservoir development process are presented (black oil model,
compositional model, etc.).
PowerPoint demonstrations illustrate some of the features of reservoir simulation
using a dataset which the student can then run on the web (with modification if
required) and plot various quantities e.g. cumulative oil, watercuts etc.
This module also contains a Glossary which the student can use for quick reference
throughout the course.

1 WHAT IS A SIMULATION MODEL?


1.1 A Simple Example of a Simulation Model
A simulation model is one which shows the main features of a real system, or
resembles it in its behaviour, but is simple enough to make calculations on. These
calculations may be analytical or numerical . By analytical we mean that the
equations that represent the model can be solved using mathematical techniques
such as those used to solve algebraic or differential equations. An analytic
solution would normally be written in terms of well know equations or
functions (x2, sin x, ex etc).
For example, suppose we wanted to describe the growth of a colony of bacteria
and we denoted the number of bacteria as N. Now if our growth model says
that the rate of increase of N with time (that is, dN/dt) is directly proportional
to N itself, then:

dN = .N
dt

(1)

where is a constant. We now want to solve this model by answering the question:
what is N as a function of time, t, which we denote by N(t), if we start with a bacterial
colony of size No. It is easy to show that, N(t) is given by:

Institute of Petroleum Engineering, Heriot-Watt University

N( t ) = N o .e . t

(2)

which is the well-known law of exponential growth. We can quickly check that
this analytical solution to our model (equation 1), is at least consistent by setting t
= 0 and noting that N = No, as required. Thus, equation 1 is our first example of a
simulation model which describes the process - bacterial growth in this case - and
equation 2 is its analytical solution. But looking further into this model, it seems
to predict that as t gets bigger, then the number N - the number of bacteria in the
colony - gets hugely bigger and, indeed, as t , the number N also . Is this
realistic ? Do colonies of bacteria get infinite in size ? Clearly, our model is not an
exact replica of a real bacterial colony since, as they grow in size, they start to use
up all the food and die off. This means that our model may need further terms to
describe the observed behaviour of a real bacterial colony. However, if we are just
interested in the early time growth of a small colony, our model may be adequate
for our purpose; that is, it may be fit-for-purpose. The real issue here is a balance
between the simplicity of our model and the use we want to make of it. This is an
important lesson for what is to come in this course and throughout your activities
trying to model real petroleum reservoirs.
In contrast to the above simple model for the growth of a bacterial colony, some
models are much more difficult to solve. In some cases, we may be able to write
down the equations for our model, but it may be impossible to solve these analytically
due to the complexity of the equations. Instead, it may be possible to approximate
these complicated equations by an equivalent numerical model. This model would
commonly involve carrying out a very large number of (locally quite simple) numerical
calculations. The task of carrying out large numbers of very repetitive calculations is
ideally suited to the capabilities of a digital computer which can do this very quickly.
As an example of a numerical model, we will return to the simple model for colony
growth in equation (1). Now, we have already shown that we have a perfectly simple
analytical solution for this model (equation 2). However, we are going to forget
this for a moment and try to solve equation 1 using a numerical method. To do this
we break the time, t, into discrete timesteps which we denote by t. So, if we have
the number of bacteria in the colony at t = 0, i.e. No, then we want to calculate the
number at time t later, then we use the new value and try to find the number at
time t later and so on. In order to do this systematically, we need an algorithm (a
mathematical name for a recipe) which is easy to develop once we have defined the
following notation:
Notation:

the value of N at the current time step n is denoted as Nn


the value of N at the next time step, n+1 is denoted as Nn+1

Clearly, it is the Nn+1 that we are trying to find. Going back to the main equation that
defines this model (equation 1), we approximate this as follows:

N n +1 N n
.N n
t

(3)

where we use the symbol, "", to indicate that equation 3 is really an approximation, or
that it is only exactly true as t 0. Equation 3 is now our (approximate) numerical

Introduction and Case Studies

model which can be rearranged as follows to find Nn+1 (which is the unknown that
we are after):

N n +1 = (1 + .t ).N n

(4)

where we have gone to the exact equality symbol, =, in equation 4 since, we are
accepting the fact that the model is not exact but we are using it anyway. This
is our numerical algorithm (or recipe) that is now very amenable to solution
using a simple calculator. More formally, the algorithm for the model would be
carried out as shown in Figure 1.
Set,

t=0

Choos e the time step size, t


Specify the initial no. of bacteria at t = 0
i.e. No and set the first value (n=0) of N n to N o
No = No

Print n, t and N (N n)

Set

N n+1 = (1 + .t). N n

Set

No

Figure 1
Example of an algorithm to
solve the simple numerical
simulation model in the
text

N n = N n+1
n = n+1
t = t + t

Time to stop ?
e.g. is t > tmax or n > nmax

Yes
End

The above example, although very simple, explains quite well several aspects of
what a simulation model is. This model is simple enough to be solved analytically.
However, it can also be formulated as an approximate numerical model which is
organised into a numerical algorithm (or recipe) which can be followed repetitively.
A simple calculator is sufficient to solve this model but, in more complex systems,
a digital computer would generally be used.

Institute of Petroleum Engineering, Heriot-Watt University

1.2 A Note on Units

Throughout this course we will use Field Units and/or SI Units, as appropriate.
Although the industry recommendation is to convert to SI Units, this makes discussion
of the field examples and cases too unnatural.
EXERCISE 1.
Return to the simple model described by equation 1. Take as input data, that we
start off with 25 bacteria in the colony. Take the value = 1.74 and take time
steps t = 0.05 in the numerical model.
(i) Using the scale on the graph below, plot the analytical solution for the
number of bacteria N(t) as a function of time between t = 0 and t = 2 (in
arbitrary time units).
(ii) Plot as points on this same plot, the numerical solution at times t = 0, 0.5, 1.0,
1.5 and 2.0. What do you notice about these ?
(iii)Using a spreadsheet, repeat the numerical calculation with a t = 0.001
and plot the same 5 points as before. What do you notice about these?

1000

N(t)
500

(i)
(ii)

Time

Introduction and Case Studies

2 WHAT IS A RESERVOIR SIMULATION MODEL?


In the previous section, we introduced the idea of a simulation model applied to
the growth of a bacterial colony. Now let us consider what we want to model - or
simulate - when we come to developing petroleum reservoirs. Clearly, petroleum
reservoirs are much more complex than our simple example since they involve many
variables (e.g. pressures, oil saturations, flows etc.) that are distributed through space
and that vary with time.
In 1953, Uren defined a petroleum reservoir as follows:
... a body of porous and permeable rock containing oil and gas through which
fluids may move toward recovery openings under the pressure existing or that
may be applied. All communicating pore space within the productive formation
is properly a part of the rock, which may include several or many individual
rock strata and may encompass bodies of impermeable and barren shale. The
lateral expanse of such a reservoir is contingent only upon the continuity of
pore space and the ability of the fluids to move through the rock pores under
the pressures available.
L.C. Uren, Petroleum Production Engineering, Oil Field Exploitation, 3rd edn.,
McGraw-Hill Book Company Inc., New York, 1953.
This fine example of old fashioned prose is not so easy on the modern ear but does
in fact say it all. And, whatever it says, then it is precisely what the modern
simulation engineer must model!

2.1 The Task of Reservoir Simulation


Let us consider the possible magnitude of the task before us when we want to model
(or simulate) the performance of a real petroleum reservoir. Figure 2 shows a
schematic of reservoir depositional system for the mid-Jurassic Linnhe and Beryl
formations in the UK sector of the North Sea. Some actual reservoir cores from
the Beryl formation are shown in Figure 3. It is evident from the cores that real
reservoirs are very heterogenous. The air permeabilities (kair) range from 1mD to
almost 3000 mD and it is evident that the permeability varies quite considerably
over quite short distances. It is common for reservoirs to be heterogeneous from
the smallest scale to the largest as is evident in these figures. These permeability
heterogeneities will certainly affect both pressures and fluid flow in the system. By
contrast, a reservoir simulation model which might be used to simulate waterflooding
in a layered system of this type is shown schematically in Figure 4. This model
is clearly hugely simplified compared with a real system. Although the task of
reservoir simulation may appear from this example to be huge, it is still one that
reservoir engineers can - and indeed must - tackle. Below, we start by listing in
general terms the activities involved in setting up a reservoir model.
One way of approaching this is to break the process down into three parts which
will all have to appear somewhere in our model:
(i) Choice and Controls: Firstly, there are the things that we have some control
over. For example:
Institute of Petroleum Engineering, Heriot-Watt University

Where the injectors and producer wells are located


The capability that we have in the well (completions & downhole equipment)
How much water or gas injection we inject and at what rate
How fast we produce the wells (drawdown)

We note that certain quantities such as injection and production rates are subject to
physical constraints imposed on us by the reservoir itself.
(ii) Reservoir Givens: Secondly, there are the givens such as the (usually very
uncertain) geology that is down there in the reservoir. There may or may not
be an active aquifer which is contributing to the reservoir drive mechanism.
We can do things to know more about the reservoir/aquifer system by carrying
out seismic surveys, drilling appraisal wells and then running wireline logs,
gathering and performing measurements on core, performing and analysing
pressure buildup or drawdown tests, etc.
(iii) Reservoir Performance Results: Thirdly, there is the observation of the results
i.e the reservoir performance. This includes well production rates of oil, water
and gas, the field average pressure, the individual well pressures and well
productivities etc.

SSW

Fluvial
mud/sand
supply
FC

OM/CS

CRS

TC
SM

SM
TC

ay

eB

rin

a
stu

OM/CS
L

Flu

lain

dp

oo

Fl
ial/

TC
TF

TF

FTD

BM

TS

SM
TF

TC

SS

SS

rrie

Ba

SM

o
Sh

a
ref

ce

SS

TCI
ETD

Fluvial/Floodplain
Facies Asociation
FC: Fluvial channel sandstones
CRS: Crevasse channel/splay sandstones
OM/L: Overbank/lake mudstone
CS: Coal swamp/marsh mudstone and coal
Estuarine Bay-Fill
Facies Association
TC: Tidal channel sandstones
TF: Lower intertidal flat sandstones
TS: Tidal shoal sandstone
SM: Salt marsh/upper intertidal flat mudstones
BM: Brackish bay mudstones
FTD: Flood tidal delta
Tidal Inlet-Barrier Shoreline
Facies Association
TCI: Tidal inlet/ebb channel sandstones
SS: Barrier shoreline sandstone
ETD: Ebb tidal delta

.15

12

km

Block diagram illustrates the gradual infilling of the


Beryl Embayment by fluvial/floodplain (Linnhe l),
estuarine-bay fill (Linnhe ll) and tidal inlet-barrier
shoreline facies sequences (Beryl Formation).
Coal
Fluvial/crevasse channel-fills
Tidal channel-fills
Tidal inlet-fills
Shoal/bars
Flood-oriented currents
Ebb-oriented currents
Longshore currents

Figure 2
Conceptual depositional
model for the Linnhe and
Beryl formations from the
middle Jurassic period (UK
sector of the North Sea).
(G. Robertson in Cores
from the Northwest
European Hydrocarbon
Provence, edited by C D
Oakman, J H Martin and
P W M Corbett, Geological
Society, London. 1997).

Introduction and Case Studies

Slab 1
Top
15855 ft

Slab 2
Top
15852 ft

Slab 3
Top
14591 ft

Slab 4
Top
14361 ft

Slab 5
Top
14358 ft

Medium-grained
Carbonate cemented
sandstone
( =14%, ka = 2mD)
- some thin clay and
carbonate rich lamination

1m

Figure 3
Cores from the midJurassic Beryl formation
from UK sector of the North
Sea. is porosity and ka is
the air permeability. (G.
Robertson in Cores from
the Northwest European
Hydrocarbon Provence,
edited by C D Oakman, J H
Martin and P W M Corbett,
Geological Society, London.
1997).

Medium-grained
ripple-laminated and
bioturbated carbonate
cemented sandstone
( =10%, ka = 1mD)

Pyritic mudstone (pm)


fine-grained bioturbated
sandstone
( =16%, ka = 29mD)

Medium to coarse-grained
cross-stratified
sandstone
( =21%, ka =1440mD)
- in fining-up units

15858 ft
Base

15855 ft
Base

14594 ft
Base

14364 ft
Base

14361 ft
Base

Coarse-grained
carbonaceous sandstone
( =20%, ka =2940mD)
- in cross-stratified,
fining-up units

Producer

Water Injector

Figure 4
A schematic diagram of a
waterflood simulation in a
3D layered model with an
8x8x5 grid. The information
which is input for a single
grid block is shown.
Contrast this simple model
with the detail in a
geological model (Figure 2)
and in the actual cores
themselves (Figure 3).

x
Inp
, ut:
cr
kx, ock,
n
S ky, k et t
og
w,
P i krw(z,
ros
c (S
S
s
w),
w)
kr
w(S
w ),

Approximate Size of Core vs. Grid Size

Institute of Petroleum Engineering, Heriot-Watt University

2.2 What Are We Trying To Do and How Complex Must Our Model Be?
Therefore, at its most complex, our task will be to incorporate all of the above features
(i) - (iii) in a complete model of the reservoir performance. But we should now
stop at this point and ask ourselves why we are doing the particular study of a given
reservoir? In other words, the level of modelling that we will carry out is directly
related to the issue or question that we are trying to address. Some engineers prefer
to put this as follows:

What decision am I trying to make?

What is the minimum level of modelling - or which tool can I use - that
allows me to adequately make that decision?

This matter is put well by Keith Coats - one of the pioneers of numerical reservoir
simulation - who said:
The tools of reservoir simulation range from the intuition and judgement of the
engineer to complex mathematical models requiring use of digital computers. The
question is not whether to simulate but rather which tool or method to use.
(Coats, 1969).
Therefore, we may choose a very simple model of the reservoir or one that is quite
complex depending on the question we are asking or the decision which we have to
make. Without giving technical details of what we mean by simple and complex,
in this context, we illustrate the general idea in Figure 5 which shows three models
of the same reservoir. The first (Figure 5a), shows the reservoir as a tank model
where we are just concerned with the gross fluid flows into and out of the system. In
Chapter 2, we will identify models such as those in Figure 5a as essentially material
balance models and will be discussed in much more detail later. The particular
advantage of material balance models is that they are very simple. They can address
questions relating to average field pressure for given quantities of oil/water/gas
production and water influx from given initial quantities and initial pressure (within
certain assumptions). However, because the material balance model is essentially
a tank model, it cannot address questions about why the pressures in two sectors
of the reservoir are different (since a single average pressure in the system is a core
assumption). The sector model in Figure 5b is somewhat more complex in that it
recognises different regions of the reservoir. This model could address the question
of different regional pressures. However, even this model may be inadequate if the
question is quite detailed such as: in my mature field with a number of active injector/
producer wells where should I locate an infill well and should it be vertical, slanted
or horizontal ? For such complicated questions, the model in Figure 5c would be
more appropriate since it is more detailed and it contains more spatial information.
This schematic sequence of models illustrates that there is no one right model for
a reservoir. The simplicity/complexity of the model should relate to the simplicity/
complexity of the question. But there is another important factor: data. It is clear
that to build models of the types shown in Figure 5, we require increasing amounts
of data as we go from Figure 5a5b5c. It is also evident that we should think
carefully before building a very detailed model of the type shown in Figure 5c, if
we have almost no data. There are some circumstances where we might build quite
10

Introduction and Case Studies

a complicated model with little data to test out hypotheses but we will not elaborate
on this issue at this point.
The simplicity/complexity of the model should relate to the simplicity/complexity of
the question, and be consistent with the amount of reliable data which we have.
(a) "Tank" Model of the Reservoir

Average Pressure
Average Saturations

Wells Offtake

=
=

P
So , Sw and Sg

Aquifer

(b) Simple Sector Model


Producer - West Flank

Producer - East Flank


Oil Leg

Aquifer

(c) Fine Grid Simulation Model of a Waterflood

Figure 5
Schematic illustrations of
reservoir models of
increasing complexity.
Each of these may be
suitable for certain types of
calculation (see text).

Injector

Producer

200ft

2000ft

We are now aware that various levels of reservoir model may be used and that the
reservoir engineer must choose the appropriate one for the task at hand. We will
assume at this point that building a numerical reservoir simulation model is the
correct approach for what we are trying to achieve. If this is so, we now address the
issue: What do we model in reservoir simulation and why do we model it ? There
are, as we have said, a range of questions which we might answer, only some of
which require a full numerical simulation model to be constructed. Let us now say
what a numerical reservoir simulation model is and what sorts of things it can (and
cannot) do.
Definition:
A numerical reservoir simulation model is a grid block model
of a petroleum reservoir where each of the blocks represents a local part of the

Institute of Petroleum Engineering, Heriot-Watt University

11

reservoir. Within a grid block the properties are uniform (porosity, permeability,
relative permeability etc.) although they may change with time as the reservoir
process progresses. Blocks are generally connected to neighbouring blocks are fluid
may flow in a block-to-block manner. The model incorporates data on the reservoir
fluids (PVT) and the reservoir description (porosities , permeabilities etc.) and their
distribution in space. Sub-models within the simulator represent and model the
injection/producer wells.
An example of numerical reservoir simulation gridded model is shown in Figure 6,
where some of the features in the above definition are evident. We now list what
needs to be done in principle to run the model and then the things which a simulator
calculate, if it has the correct data.
To run a reservoir simulation model, you must:
(a) Gather and input the fluid and rock (reservoir description) data as outlined above;
(b) Choose certain numerical features of the grid (number of grid blocks, time
step sizes etc);
(c) Set up the correct field well controls (injection rates, bottom hole pressure
constraints etc.); it is these which drive the model;
(d) Choose which output (from a vast range of possibilities) you would like to have
printed to file which you can then plot later or - in some cases - while the
simulation is still running.
The output can include the following (non-exhaustive) list of quantities:

The average field pressure as a function of time


The total field cumulative oil, water and gas production profiles with time
The total field daily (weekly, monthly, annual) production rates of each
phase: oil, water and gas
The individual well pressures (bottom hole or, through lift curves, wellhead)
over time
The individual well cumulative and daily flowrates of oil, water and gas
with time
Either full field or individual well watercuts, GORs, O/W ratios with time
The spatial distribution of oil, water and gas saturations throughout the
reservoir as functions of time i.e. So(x,y,z;t), Sw(x,y,z;t) and Sg(x,y,z;t)

Some of the above quantities are shown in simulator output in Figure 7. This field
example is for a Middle East carbonate reservoir where the structural map is shown in
Figure 7(d). Figure 7(a) shows the field and simulation results for total oil and water
cumulative production over 35 years of field life. Figure 7(b) shows the actual and
modelled average field pressure. The type of results shown in Figures 7(a) and 7(b)
are very common but the modelling of the RFT (Repeat Formation Tester) pressure
shown in Figure 7(c) is less common. The RFT tool measures the local pressure at a
given vertical depth and, in this case, it can be seen that the reservoir comprises of
three zones each of ~ 100 ft thick and each is at a different pressure. This indicates that
12

Introduction and Case Studies

pressure barriers exist (i.e. flow is restricted between these layers). This is correctly
modelled in the simulation. This is an interesting and useful example of how reservoir
simulation is used in practice.

Figure 6
An example of a 3D
numerical reservoir
simulation model. The
distorted 3D grid covers
the crestal reservoir and a
large part of the aquifer
which is shown dipping
down towards the reader.
Oil is shown in red and
water is blue and a vertical
projection of a cross-section
at the crest of the reservoir
is shown on the x/z and
y/z planes on the sides of
the perspective box. Two
injectors can be seen in the
aquifer as well as a crestal
horizontal well. Two faults
can be seen at the front
of the reservoir before the
structure dips down into the
aquifer. The model contains
25,743 grid blocks.

Note that a vast quantity of output can be output and plotted up and the post-processing
facilities in a reservoir simulator suite of software are very important. There is no
point is doing a massively complex calculation on a large reservoir system with
millions of grid blocks if the output is so huge and complex that it overwhelms the
reservoir engineers ability to analyse and make sense of the output. In recent years,
data visualisation techniques have played on important role in analysing the results
from large reservoir simulations.

600
500

Observed Oil

400

Modelled Oil

300
200

Observed Water
Modelled Water

100
0

10

15

20

25

30

35

Year of Production
(a) Full field history match of cumulative oil and water production
3500
Institute of Petroleum Engineering, Heriot-Watt University
ure (psia)

Figure 7 (a) to (d)


Example of some typical
reservoir simulator output.
From SPE36540,
Reservoir Modelling and
Simulation of a Middle
Eastern Carbonate
Reservoir, M.J. Sibley,
J.V. Bent and D.W. Davis
(Texaco), 1996.

Cumulative Production (MMB)

700

3000

13

Cumulative Pro

Modelled Oil
300
200

Observed Water
Modelled Water

100
0

10

15

20

25

30

35

Year of Production

Average Pressure (psia)

(a) Full field history match of cumulative oil and water production
3500
3000
2500

Observed Data
Modelled Data

2000
1500

10

15

20

25

30

35

Year of Production

Figure 7b

(b) Full field history match of volume weighted pressure

Datum

Observed
Modelled
Depth (ft.)

-100

-200

-300
1000

1500

1000

2500

3000

(c) Match of RFT pressure data by reservoir simulation model at Year 30

Figure 7c

A Lower Cretaceous

Carbonate Reservoir in the


Arabian Peninsula

Most wells drilled in 1955-1962

> 600 MMBO produced by


early 1980s

-this study 1992


C

Drilled
New Location
Injector Location

1 Mile

(d) Field structural map with 50' contour interval

14

Convert to Injector

Figure 7d

Introduction and Case Studies

How some of this output might be used is illustrated schematically in Figure 8.


This is an imaginary case where the reservoir study is to consider the best of four
options in Field A: Option 1 - to continue as present with the waterflood; Option 2
- upgrade peripheral injection wells; Option 3 - upgrade injectors and drill six new
injectors; Option 4 - drill four new infill wells. Clearly, it is much cheaper to model
these four cases than to actually do one of them. The important quantities are the
oil recovery profiles for each case compared with the scenario where we simple
proceed with the current reservoir development strategy (Option 1). Of course, we
do not know whether the forward predictions which we are taking as what would
happen anyway, are actually correct. Likewise, we may be unsure of how accurate
our forward predictions are for each of the various scenarios. In fact, an important
aspect of reservoir simulation is to assess each of the various uncertainties which
are associated with our model. This would ideally lead to range of profiles for any
forward modeling but we will deal with this in detail later. We discuss the handling
of uncertainties in rather more detail in Section 3.8. of this Chapter.
In the schematic case shown in Figures 8(a) - 8(g) we note that:
(i) The areal plan of the reservoir is given showing injector and producer well
location in Figure 8(a);
(ii) The corresponding stratification/lithology of the field is shown along the well
A-B-C-D transect in Figure 8(b);
(iii) Figures 8(c) and 8(d) show the areal grid and the vertical grid, respectively;
(iv) The forward predictions of cumulative oil for the various options are shown
in Figure 8(f). Note that Option 3 produces most oil (but it involves drilling
six additional injection wells);
(v) The economic evolution of each option using the predicted oil recovery profiles
in Figure 8(f) is shown in Figure 8(g) (where NPV = Net Present Value; IRR =
Interval Rate of Return: these are economic measures explained in the economics
module of the Heriot-Watt distance learning course). Note that option 4 emerges
in the most economic case although it produces rather less oil than option 3.
Injector
Producer
A
C

Figure 8
Schematic example of how
reservoir simulation might
be used in a study of four
field development options
(see text).

(a) Field A areal plan showing injector and producer well locations; lithology is
given from wells A, B, C and D

Institute of Petroleum Engineering, Heriot-Watt University

15

Sand 1

Sand 2
Sand 3
Sand 4

Figure 8 (b)

(b) Schematic vertical cross-section showing the lithology across the field through
4 wells A, B, C and D
A

A
B

Figure 8 (c)

(c) Reservoir simulation (areal) grid showing current well locations.

A
A
B

NZ = 8
NZ = 8

D
D

(d) Reservoir simulation vertical cross-sectional grid showing current well locations.

16

Figure 8 (d)

Introduction and Case Studies

The grid has 8 blocks in the z- direction representing the thickness of the
formation as shown below; NZ = 8. Note that the vertical grid is not uniform.
Periferal Injectors
A

Periferal Injectors
C
Periferal Injectors

B
B

D
D

Infill Wells

(e) Option 1- continue as at present; Option 2 - upgrade peripheral injection wells;


Option 3- upgrade injectors + add 6 new injectors; Option 4 - drill four new infill
wells.
Option 4
Infill Wells
Option 3

Infill Wells

Option 4
Option 3
Option 4

Cumulative
Oil Oil
Cumulative

Cumulative Oil

Figure 8 (e)

Option 3
Continue as at present (do nothing) Option 1
Option 2

Time
Continue as at present (do nothing) Option 1
Option 2
Continue as at present (do nothing) Option 1

Time

Option 2

Time

Figure 8 (f)

Simulated oil recovery results for


various4options
3
1

NPV NPV
or IRR
or IRR

NPV or IRR

(f)

3
3

1
1

2
2

4
4

Option

Option
Option

Figure 8 (g)
Institute of Petroleum Engineering, Heriot-Watt University

17

(g) Economic evaluation of options - NPV or IRR


Now consider what we are actually trying to do in a typical full field reservoir simulation
study. There is a short answer to this is often said in one form or another: it is that the
central objective of reservoir simulation is to produce future predictions (the output
quantities listed above) that will allow us to optimise reservoir performance. At the
grander scale, what is meant by optimise reservoir performance is to develop the
reservoir in the manner that brings the maximum economic benefit to the company.
Reservoir simulation may be used in many smaller ways to decide on various
technical matters although even these - for example the issue illustrated in Figure 8
- are usually reduced to economic calculations and decisions in the final analysis as
indicated in Figure 8(g).

3 FIELD APPLICATION OF RESERVOIR SIMULATION


3.1 Reservoir Simulation at Appraisal and in Mature Fields

Up to this point, we have considered what a numerical reservoir simulation model


is and we have touched on some of the sorts of things that can be calculated. Rather
than continue with a discussion of the various technical aspects of reservoir simulation
one by one, we will proceed to three field applications of reservoir simulation. These
studies will raise virtually all of the technical terms and concepts and many of the
issues that will be studied in more detail later in this course. The important terms
and concepts will be italicised and will appear in the Glossary at the front of this
chapter.
Reservoir simulation may be applied either at the appraisal stage of a field
development or at any stage in the early, middle or late field lifetime. There are clearly
differences in what we might want to get out of a study carried out at the appraisal
stage of a reservoir and a study carried out on a mature field.
Appraisal stage: at this stage, reservoir simulation will be a tool that can be used to
design the overall field development plan in terms of the following issues:

The nature of the reservoir recovery plan e.g. natural depletion, waterflooding,
gas injection etc.

The nature of the facility required to develop the field e.g. a platform, a subsea
development tied back to an existing platform or a Floating Production System
(for an offshore fileld).

The nature and capacities of plant sub-facilities such as compressors for


injection, oil/water/gas separation capability.

The number, locations and types of well (vertical, slanted or horizontal) to be


drilled in the field.

The sequencing of the well drilling program and the topside facilites.

18

Introduction and Case Studies

It is during the initial appraisal stage that many of the biggest - i.e. most expensive
- investment decisions are made e.g. the type of platform and facilities etc. Therefore,
it is the most helpful time to have accurate forward predictions of the reservoir
performance. But, it is at this time when we have the least amount of data and,
of course, very little or no field performance history (there may be some extended
production well tests). Therefore, it seem that reservoir simulation has a built-in
weakness in its usefulness; just when it can be at its most useful during appraisal is
precisely when it has the least data to work on and hence it will usually make the
poorest forward predictions. So, does reservoir simulation let us down just when we
need it most? Perhaps. However, even during appraisal, reservoir simulation can
take us forward with the best current view of the reservoir that we have at that time,
although this view may be highly uncertain. As we have already noted, if major
features of the reservoir model (e.g. the stock tank oil initially in place, STOIIP) are
uncertain, then the forward predictions may be very inaccurate. In such cases, we
may still be able to build a range of possible reservoir models, or reservoir scenarios,
that incorporate the major uncertainties in terms of reservoir size (STOIIP), main
fault blocks, strength of aquifer, reservoir connectivity, etc. By running forward
predictions on this range of cases, we can generate a spread of predicted future field
performance cases as shown schematically in Figure 9. How to estimate which of
these predictions is the most likely and what the magnitude of the true uncertainties
are is very difficult and will be discussed later in the course.

Cumulative Oil Recovery (STB)

"Optimistic" Case

Figure 9
Spread of future predicted
field performances from a
range of scenarios of the
reservoir at appraisal.

2005

"Pessimistic" Case
Most Probable Case

2010

Time (Year)

2015

For example, scenarios for various cases may involve:

Different assumptions about the original oil in place (STOIIP; Stock Tank Oil
Originally In Place).

Different values of the reservoir parameters such as permeability, porosity,


net-to-gross ratio, the effect of an aquifer, etc..

Major changes in the structural geology or sedimentology of the reservoir


e.g. sealing vs. leaky faults in the system, the presence/absence of major
fluvial channels, the distribution of shales in the reservoir etc..

Institute of Petroleum Engineering, Heriot-Watt University

19

Mature field development: we define this stage of field development for our purposes
as when the field is in mid-life; i.e. it has been in production for some time
(2 - 20+ years) but there is still a reasonably long lifespan ahead for the field, say
3 - 10+years. At this stage, reservoir simulation is a tool for reservoir management
which allows the reservoir engineer to plan and evaluate future development options
for the reservoir. This is a process that can be done on a continually updated basis.
The main difference between this stage and appraisal is that the engineer now has
some field production history, such as pressures, cumulative oil, watercuts and GORs
(both field-wide and for individual wells), in addition to having some idea of which
wells are in communication and possibly some production logs. The initial reservoir
simulation model for the field has probably been found to be wrong, in that it fails
in some aspects of its predictions of reservoir performance e.g. it failed to predict
water breakthough in our waterflood (usually, although not always, injected water
arrives at oil producers before it is expected). By the way, if the original model
does turn out to be wrong, this does not invalidate doing reservoir simulation in the
first place. (Why do you think this is so?)
At this development stage, typical reservoir simulation activities are as follows:

Carrying out a history match of the (now available) field production history
in order to obtain a better tuned reservoir model to use for future field performance
prediction

Using the history match to re-visit the field development strategy in terms
of changing the development plan e.g. infill drilling, adding extra injection
water capability, changing to gas injection or some other IOR scheme etc.

Deciding between smaller project options such as drilling an attic horizontal


well vs. working over 2 or 3 existing vertical/slanted wells

It may be necessary to review the equity stake of various partner companies


in the field after some period of production although this typically involves
a complete review of the engineering, geological and petrophysical data prior
to a new simulation study

The reservoir recovery mechanisms can be reviewed using a carefully history


matched simulation model e.g. if we find that, to match the history, we must
reduce the vertical flows (by lowering the vertical transmissibility), we may
wish to determine the importance of gravity in the reservoir recovery mechanism.
(Coats (1972) refers to this as the educational value of simulation models
and it is a part of good reservoir management that the engineer has a good
grasp of the important reservoir physics of their asset.)

There are many reported studies in the SPE literature where the simulation model is
re-built in early-/mid-life of the reservoir and different future development options
are assessed (e.g. see SPE10022 attached to this chapter).
Late field development: we define this stage of field development as the closing few
years of field production before abandonment. A question arises here as to whether
the field is of sufficient economic importance to merit a simulation study at this stage.
20

Introduction and Case Studies

A company may make the call that it is simply not worth studying any further since
the payback would be too low. However there are two reasons why we may want
to launch a simulation study late in a fields lifetime. Firstly, we may think that,
although it is in far decline, we can develop a new development strategy that will
give the field a new lease of life and keep it going economically for a few more
years. For example, we may apply a novel cheap drilling technology, or a program
of successful well stimulation (to remove a production impairment such as mineral
scale) or we may wish to try an economic Improved Oil Recovery (IOR) technique.
Secondly, the cost of field abandonment may be so high - e.g. we may have to remove
an offshore structure - that almost anything we do to extend field life and avoid this
expense will be economic. This may justify a late life simulation study. However,
there are no general rules here since it depends on the local technical and economic
factors which course of action a company will follow. In some countries there may
be legislation (or regulations) that require that an oil company produces reservoir
simulation calcualtions as part of their ongoing reservoir management.

3.2 Introduction to the Field Cases

Three field cases are now presented. We reproduce the full SPE papers describing
each of these reported cases. In the text of each of these papers there are margin
numbers which refer to the Study Notes following the paper. We use these to explain
the concepts of reservoir simulation as they arise naturally in the description of a field
application. In fact, you may very well understand many of the term immediately
from the context of their description in the SPE paper.
The three field examples are as follows:
Case 1: The Role of Numerical Simulation in Reservoir Management of a West
Texas Carbonate Reservoir, SPE10022, presented at the International Petroleum
Exhibition and Technical Symposium of the SPE, Beijing, China, 18 - 26 March
1982, by K J Harpole and C L Hearn.
Case 2: Anguille Marine, a Deepsea-Fan Reservoir Offshore Gabon: From Geology
Toward History Matching Through Stochastic Modelling, SPE25006, presented at the
SPE European Petroleum Conference (Europec92), Cannes, France, 16-18 November
1992, by C.S. Giudicelli, G.J. Massonat and F.G. Alabert (Elf Aquitaine)
Case 3: The Ubit Field Rejuvenation: A Case History of Reservoir Management of
a Giant Oilfield Offshore Nigeria, SPE49165, presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, LA, 27-30 September 1998, by C.A. Clayton
et al (Mobil and Department of Petroleum Resources, Nigeria)
These cases were chosen for the following main reasons:

They are all good technical studies that illustrate typical uses of reservoir
simulation as a tool in reservoir management (we have deliberately taken all
cases at the middle and the mature stages of field development since much
more data is available at that time);

They introduce virtually all of the main ideas and concepts of reservoir
simulation in the context of a worked field application. As these concepts

Institute of Petroleum Engineering, Heriot-Watt University

21

and specialised terms arise, they are explained briefly in the study notes although
more detailed discussion will appear later in the course. Compact definitions
of the various terms are given in the Glossary at the front of this module;

They are all well-written and use little or no mathematics;

By choosing an example from the early 1980s, the early/mid 1990s and the late
1990s, we can illustrate some of the advances in applied reservoir simulation
that have taken place over that period (this is due to the availability of greater
computer processing power and also the adoption of new ideas in areas such
as geostatistics and reservoir description).

How you should read the next part of the module is as follows:

Read right through the SPE paper and just pay particular attention when there
is a Study Note number in the margin;

Go back through the paper but stop at each of the Study Notes and read
through the actual point being made in that note.

As noted above, all the main concepts that are introduced can also be found
in the Glossary which should be used for quick reference throughout the
course or until you are quite familiar with the various terms and concepts in
reservoir simulation.
See SPE 10022 paper in Appendix

3.3 Case 1: The West Seminole Field Simulation Study (SPE10022,


1982)

Case 1:
The Role of Numerical Simulation in Reservoir Management of a West
Texas Carbonate Reservoir, SPE10022, presented at the International Petroleum
Exhibition and Technical Symposium of the SPE, Beijing, China, 18 - 26 March
1982. by K J Harpole and C L Hearn.
Summary: This paper presents a study from the early 1980s where a range of reappraisal strategies for a mature carbonate field are being evaluated using reservoir
simulation. For example, possible development strategies include the blowdown
of the gas cap or infill drilling. They explicitly state in their opening remarks that
their central objective is to optimise reservoir performance by choosing a future
development strategy from a range of defined options. The structure of the study is
very typical of the work flow of a field simulation study, viz Introduction; Reservoir
Description; Simulation Model; History Matching; Future Performance; Conclusions
and recommendations. Although this paper is almost 20 years old, it introduces the
reader in a very clear way to virtually all the concepts of conventional reservoir
simulation.
Location maps and general reservoir structure shown in Figures 1 and 2 of SPE
10022.

22

Introduction and Case Studies

Study Notes Case 1:

1. States explicitly that the objective of the study is to optimise reservoir performance
as discussed in the introductory part of this module.
2. Raises the issue of an accurate reservoir description being required and this is
emphasised throughout this paper.
3. An interesting point is raised comparing the carbonate reservoir of this study
broadly to sandstone reservoirs. It notes that the post-depositional diagenetic effects
are of major importance in the West Seminole field in that they affect the reservoir
continuity and quality i.e. the local porosity and permeability. In contrast, it is noted
that sandstone reservoir are mainly controlled by their depositional environment
and tend to show less diagenetic overprint. However, a point to note is that the
broad outline and work flow of a numerical reservoir simulation study are quite
similar for both carbonate and sandstone reservoirs.
4. Carbonate diagenetic processes include dolomitisation (dolomite = CaMg(CO3)2),
recrystallisation, cementations and leaching. This geochemical information is not
directly used in the simulation model but it is important since it leads to identification
of reservoir layer to layer flow barriers (see below).
5. Strategy: Previous gas re-injection into the cap + peripheral water injection =>
not very successful. They want to implement a 40 acre, 5-spot water flood; see
Fig. 3. A 5-spot is a particular example of a pattern flood which is appropriate
mainly for onshore reservoirs where many wells can be drilled with relatively close
spacing (see Waterflood Patterns in the Glossary).
6a. They raise the issue of vertical communication between the oil and gas zones.
This is an excellent example of an uncertain reservoir feature that can be modelling
using a range of scenarios from free flow between layers to zero interlayer flow + all
cases in between. Therefore, we can run simulations of all these cases and see which
one agrees best with the field observations (which is what they do, in fact).
6b. The vertical communication - or lack of it - will affect flow between the oil and
gas zones which may lead to loss of oil to the gas cap; see Figure 4.
7. States the structure of the simulation study work flow: Accurate reservoir description
- Develop the simulation model (perform the history match - see below - use model
for future predictions - evaluate alternative operating plans). A history match is when
we adjust the parameters in the simulation model to make the simulated production
history agree with the actual field performance (expanded on below).
8a. A lengthy geological description of the reservoir is given where the depositional
environment is described - reference is made to extensive core data (~7500 ft. of
core).
8b. The impact of the geology/diagenesis in the simulation model is discussed here.
There is evidence of field wide barriers due to cementation with anhydrite which may
reduce vertical flows. This is important since it gives a sound geological interpretation
to the existence of the vertical flow barriers. Therefore, if we need to include this to
Institute of Petroleum Engineering, Heriot-Watt University

23

match the field performance, we have some justification or explanation for it rather
than it simply being a fiddle factor in the model.
9. Figure 5 shows the 6 major reservoir layers where the interfaces between the layers
are low , low k anhydrite cement zones. Again, these may be explained from the
depositional environment and the subsequent diagenetic history of the reservoir.
10. 7500 ft. of whole core analysis for the W. Seminole field was available which
was digitised for computer analysis (not common at that time, late 1970s). This is
very valuable information and is often not available.
11. Permeability distributions in the reservoir are shown in Fig. 6 and these data
are vital for reservoir simulation. Dake (1994; p.19) comments on this type of data:
What matters in viewing core data is the all-important permeability distribution
across the producing formations; it is this, more than anything else, that dictates the
efficiency of the displacement process.
12. They note that no consistent k/ correlation is found in this system (which is
quite common in carbonates). Often some approximate k/ correlation can be found
for sandstones (e.g. see k/ Correlations in the Glossary).
13. The W. Seminole field does exhibit a distinctly layered structure and the
corresponding permeability stratification in the model is shown in Fig. 7.
14a. Pressure transient work - again gives important ancillary information on
the reservoir. The objectives of this work were to determine whether there was
(i) directional permeability effects, directional fracturing or channelling; (ii) the
degree of stratification in the reservoir; (iii) evaluation of the pay continuity
between the injectors and producers
14b. No evidence of channelling or obvious fracture flow system
14c. Distinct evidence was seen for: (a) the presence of a layered system; (b) restricted
communication between layers (P 200 - 250 psi between layers). This is vital
information since it gives an immediate clue that there is probably not completely free
flow between layers i.e. there are barriers to flow as suspected from the geology.
14d. Finally on this issue, there is pressure evidence of arithmetically averaged
permeabilities. This is again typical of layered systems.
15. Native state core tests are referred to from which they obtained steady-state
relative permeabilities. These could be very valuable results but no details given
here. NB it appears that only one native state core relative permeability was actually
measured. This is probably too little data but reflects the reality in many practical
reservoir studies that often the engineer does not have important information; however,
we just have to get on with it.
16. In this study the reservoir simulator which they used was a commercial Black Oil
Model (3D, 3 phase - oil/water/gas). Modelling carried out on the main dome portion
of the reservoir. This is done quite often in order to simplify the model and to focus
24

Introduction and Case Studies

on the region of the field of interest (and importance in terms of oil production). A
no flow boundary is assumed in the model on the saddle with the east dome (justified
by different pressure history). Again, this is supported by field evidence but it may
also be a simplifying judgement to avoid unnecessary complication in the model.
17a. The grid structure used in the simulations is shown in Fig 8. The particular grid
that is chosen is very important in reservoir simulation. An areal grid of 288 blocks
( 16 x 18 blocks) - about 10 acre each is taken along with six layers in the vertical
direction; i.e. a total of 1728 blocks. This would be a very small model by todays
standards and could easily be run on a PC - this was not the case in late 1970s.
17b. They refer to changing the transmissibilities between grid blocks in order to
reduce flows. (See Glossary for exact definition of transmissibility.)
18. The following three concepts are closely related (see Pseudo-isation and
Upscaling in the Glossary):
18a. Grid size sensitivity: Refers to the introduction of errors due to the coarsness
of the grid known as numerical dispersion.
18b. The very important concept of pseudo--relative permeability is introduced here
(Kyte and Berry, 1975). Pseudos are introduced in order to control numerical
dispersion and account for layering. In essence, the use of pseudos can be seen as
a fix up for using a coarse grid structure.
18c. Corresponding coarse and fine grid reservoir models are shown in Fig. 9.
They note that the fine grid model uses rock relative permeabilities while the coarse
grid model uses pseudo relative permeabilities.
19. History Matching: The basic idea of history matching is that the model input
is adjusted to match the field pressures and production history. This procedure is
intended as being a way of systematically adjusting the model to agree with field
observations. Hopefully we can change the correct variables in the model to get
a match e.g. we may examine the sensitivity to changes in vertical flow barriers in
order to find which level of vertical flow agrees best with the field (indeed, this is
done in this study). See History Matching in the Glossary.
20a. Early mechanism identified as solution gas drive and assistance from expansion.
Some initial discussion of field experience and numerical simulation conclusions is
presented and developed in these points.
20b. They note some problems with data from early field life. (i) Complicated by
free gas production; (ii) channelling due to poor well completions; (ii) no accurate
records on gas production for the first 6 years.
20c. The actual field history match indicates that approx. 8 - 10 BCF of gas must have
been produced over this early period in order to match the field pressures. This is a
use of a material balance approach in order to find the actual early STOIIP (STOIIP
= Stock Tank Oil Initially In Place).

Institute of Petroleum Engineering, Heriot-Watt University

25

21a. They present a description of some adjustments to the history match - but overall
it is very good (which they attribute to extensive core data).
21b. Some highlighting of problems with earlier water injection .
21c. The actual history match of reservoir pressure and production is shown in Fig.
10. This is a good history match but think of which field observable - gas production,
water production or average field pressure - is the easiest/most difficult to match?
22. A good description of their study of the sensitivity to vertical communication
is given at this point. This is examined by adjusting the vertical transmissibilities.
They look at the following cases: (i) no barriers; (ii) moderate barrier; (iii)
strong barriers and (iv) no-flow barriers. Most of the sensitivties are for the
moderate and strong barrier cases.
23a. Results showed that => strong barrier case is best but some problem high
GOR wells are encountered randomly spaced through the field. They diagnosed
and simulated this as behind the pipe gas flow in these wells to explain the
anomalies in the field observations. This is quite a common explanation that
appears in many places.
23b. Layer differential pressures up to 200 - 250 psi can only be reproduced for
the strong barrier case. In simulation terms, this is probably the strongest evidence
that this is the best case match.
24. The strong barrier case was chosen as the base case and this was used for
the predictive runs. The base case predictions refer to the cases which essentially
continue the current operations and these are shown in Fig. 11.
25. The strategies looked at for the future sensitivities are listed as follows: (i) change
rate of water injection; (ii) management of gas cap voidage i.e. increase of gas and
blowdown at different times; (iii) infill drilling.
26a. Outlines the problems/issues for various strategies as follows: (i) shows vertical
communication is very importance - it has a major impact on predicted reservoir
performance; (ii) shows that can avoid high future P between gas cap and oil
zone by high water injection or early blowdown; (iii) shows better development
strategy is to keep low P e.g. increase gas injection or infill drill. Finally, shows
infill drilling is the most attractive option and the forward prediction for this case is
shown in Figure 12.
26b. Table showing some alternatives in text.
27a. A brief summary of the best future development option is given which is: (i)
infill drilling as the best option; (ii) water injection increased concurrently with the
drilling program to maintain voidage replacement (but prevent the over-injection of
water).
27b. For completeness, it is explained why other plans are not as attractive; i.e.
blowdown of gas cap before peak in waterflood production rate would significantly
reduce oil recovery.
26

Introduction and Case Studies

28. A reasonably good initial forward prediction from 1978 - 1981 is shown in Figs.
13 and 14.
29. Conclusions are given which, in summary, are as follows:
1.

Detailed reservoir description essential for numerical modelling.

2.

Carbonate - both primary and post-depositional diagenetic factors are


important.

3.

Waterflood in W. Seminole very sensitive to vertical permeability.

4.

Vertical permeability is broadly characterised using 3D numerical


simulation.

5.

Understanding of reservoir response (mechanism) essential to good


management.

6.

Management of W. Seminole field best if minimum _P between oil zone and


gas cap (lower losses of oil --> gas cap) by: (i) infill drilling; (ii) controlling
water injection rates to maintain voidage replacement - dont over-inject; (iii)
careful management of voidage replacement into gas cap.

Important terms and concepts introduced in SPE10022:


Specific to Reservoir Simulation: history match, permeability distribution, black
oil model, grid structure, transmissibility, numerical dispersion, pseudo--relative
permeabilites.
General terms: 5-spot water flood, permeability distribution, k/ correlation, steadystate relative permeability, rock relative permeabilities, solution gas drive, material
balance, infill drilling, voidage replacement.

3.4 Ten Years Later - 1992

An interesting snapshot of where reservoir simulation technology had reached 10


years after the West Seminole study can be seen in the following papers:
From the proceedings of the SPE 67th Annual Technical Conference, Washington,
DC, 4-7 October 1992:
SPE24890: From Stochastic Geological Description to Production Forecasting in
Heterogeneous Layered Systems, K. Hove, G. Olsen, S. Nilsson, M. Tonnesen and
A. Hatloy (Norsk Hydro and Geomatic)
Summary: This paper describes the transfer of data from a detailed gridded
stochastic geological model to a more coarsely gridded reservoir simulation model.
It is essentially a field application of a methodology described in a previous paper
from the same company (Damsleth et al, 1992; Damsleth, E., Tjolsen, C.B., Omre,
H. and Haldonsen, H.H., A Two Stage Stochastic Model Applied to a North Sea
Reservoir, J. Pet. Tech., pp. 402-408, April 1992). The two step procedure involves
a first step of constructing the geological architecture of the reservoir followed by a
Institute of Petroleum Engineering, Heriot-Watt University

27

second stage where the petrophysical values are assigned to each building block in the
geological model. The consequences of making various assumptions in the gridding
are evaluated for water, gas and WAG (water-alternating-gas) injection. They note
that is it very important to represent the main geological features in the gridded model.
It was also noted that, when a regular coarse grid was used, the contrast in properties
of this heterogeneous reservoir were smoothed out by the averaging process and
in most cases led to a more optimistic predicted production performance. That is,
the more stochastic models led to a reduction in predicted recovery compared with
conventional coarse gridded models.
In the proceedings of the SPE European Petroleum Conference, Cannes, France,
16-18 November 1992. A session at this conference produced the following
selection of reservoir simulation papers:
SPE25008: Reservoir Management of the Oseberg Field After Four Years,
S. Fantoft (Norsk Hydro)
Summary: The Oseberg Field (500x106 Sm3 oil; 60x106 Sm3 gas) comprises of
seven partly communicating reservoirs. Both water and gas are being injected and
modelled in this study and results indicate over 60% recovery in the main reservoir
units. The modelling results indicate that the plateau production will be extended
by the use of horizontal wells. The objective of the simulation study was exactly
this - i.e. to maximise the plateau and improve ultimate oil recovery. This is a very
competent simulation study and - although details are not given - it is stated that
the geological model is updated annually based on information from new wells.
It establishes several aspects of the reservoir mechanics and makes a number of
recommendations for operating practice in the future. In other respects, this is quite
a conventional study.
SPE25057: The Construction and Validation of a Numerical Model of a Reservoir
Consisting of Meandering Channels, W. van Vark, A.H.M. Paardekam, J.F. Brint
J.B. van Lieshout and P.M. George (Shell)
Summary: This study focuses on a reservoir which has low sandbody connectivity and
which is interpreted as a meandering channel fluvial system. Two years of depletion
data is available and one of the aims of the study was to evaluate the possibility of
performing a waterflood in this field. They identified a problem in that the sandbody
connectivity was lower than might be expected from the sedimentology alone and it
was conjectured that this might be due to minor faulting with throws of a few meters.
This study again emphasises the importance of the reservoir geology and tries to
relate the performance back to this. The geological model is also an early practical
example of using a voxel representation of the system - approx. 128,000 voxels
were used in the model. They noted that the original (sedimentological) models
gave over optimistic connectivity. An acceptable match to observed field pressures
by including some level of smaller scale faulting.
SPE25059: Development Planning in a Complex Reservoir: Magnus Field UKCS
Lower Kimmeridge Clay Formation (LKCF), A.J. Leonard, A.E. Duncan, D.A.
Johnson and R.B. Murray (BP Exploration Operating Co.)
Summary: This simulation study was carried out on the geologically complex,
low net to gross LKCF (rather than on higher net to gross Magnus sands studied
28

Introduction and Case Studies

previously). The objective was to formulate a development plan for the LKCF which
would accelerate production from these sands. Stochastic modelling techniques
were integrated into more conventional deterministic models and various options
were screened for inherent uncertainty and risks. The study concluded that a phased
water injection scheme was the best way forward with the phasing being used to
manage and offset the considerable geological risks. Ranges of expected recovery
were generated and an incremental recovery of 60 MMstb was predicted increasing
the total reserve of the LKCF by a factor of x2.4. This study also demonstrated the
importance of inter-disciplinary team work to overcome the previously inhibiting
high risks involved.
The proceedings of Europec92 also included the following paper:
SPE25006: Anguille Marine, a Deepsea-Fan Reservoir Offshore Gabon: From
geology Toward History Matching Through Stochastic Modelling, C.S. Giudicelli,
G.J. Massonat and F.G. Alabert (Elf Aquitaine)
This paper is such a good example of contemporary studies at that time, that
this is chosen as our Case 2 example and is presented in some detail in the
next section.

3.5 Case 2: The Anguille Marine Study (SPE25006,1992)

Case 2:
Anguille Marine, a Deepsea-Fan Reservoir Offshore Gabon: From
geology Toward History Matching Through Stochastic Modelling, SPE25006,
presented at the SPE European Petroleum Conference (Europec92), Cannes, France,
16-18 November 1992, by C.S. Giudicelli, G.J. Massonat and F.G. Alabert (Elf
Aquitaine)
See SPE 25006 paper in Appendix
Summary: The Anguille Marine Field in Gabon has 25 years of production history.
The waterflood performance indicated severe sedimentary heterogeneity as the field
is known to have been deposited in a deep water fan sedimentary environment. This
paper is one of the first to refer to the multi-scale nature of the heterogeneity (5 scales
were studied) and to refer this back to the sequence stratigraphy of the depositional
environment. The sequence stratigraphic approach allowed the field to be divided into
the main types of turbiditic geometries (channels, lobes, slumps, laminated facies).
Fine scale models (> 2 million grid blocks) were generated using geostatistical
techniques and several issues were raised concerning both the geological model
and the upscaling process itself. This is a very good example of an early integrated
geology(sedimentology)/engineering study in reservoir simulation. The multi-scale
nature of the heterogeneity is well related back to the geology.

Study Notes Case 2:

1. Depositional environment: the Anguille Marine field is a deep sea fan environment
(i.e a turbidite) with a low sand/shale ratio. This geological description opens the
discussion (unusual for previous simulation studies) and the geology features heavily in
the flow properties and hence in the geological and reservoir models of this field.
2. Sequence stratigraphy: A more modern feature of reservoir simulation is that the
five identified scales of heterogeneity are recognised and some attempt is made to
Institute of Petroleum Engineering, Heriot-Watt University

29

incorporate them into the 3D simulation model. These scales are also firmly linked
to the geology (sedimentology) through the principles of sequence stratigraphy.
3. Geostatistics: Reference is made to how the geological features constrain the
fine scale 3D models (of > 2 million blocks - which was large for the time) using
geostatistical techniques. By the early 90s, the use of geostatistical methods was
becoming more widespread and how it has been applied in this case is covered better
by Refs. 1 - 5 in this paper.
Location and structure maps of Anguille Marine are given in Figures 1 - 4.
4. Brief field facts: Discovery 1962; primary depletion commenced in 1966 but reservoir
pressure fell rapidly over the next 2 - 3 years and GOR increased; waterflooding from
1971 restored pressure support but channelling led to early water breakthrough; infill
drilling not very successful due to lack of current understanding of complex reservoir
geology; new approach in 1990 focused more strongly on the reservoir geology of this
heterogeneous low sand/shale ratio system recognising the characteristic geometries
of a tubditic fan - lobes channels, levees, slumps, laminated facies etc.
5. The approach: It is important in all reservoir simulation studies to have a clear
logic to how we approach the simulation of a large complex reservoir system. Here
they describe their general methodology although details are in Refs. 1 - 4 at the end
of the paper. Basically they: describe and model upper reservoir/ extend to the whole
reservoir/ try to translate the geological model to a practical simulation model. On
the latter issue they describe the use of partial models where just a smaller sector
of the reservoir is studied but lessons are taken back into the full model.
6. Reservoir description: Section 2 of the paper gives a sedimentological description
of the reservoir as a slope-apron fan of complex lithology (depositional model Figure
3) in which 14 (simplified) facies were retained; criteria of composite log recognition of
various facies shown in Figure 5. Some contradictory water breakthrough observations
were noted. Table 1 gives sedimentary body dimensions (lengths and widths) for
channels, lobes. levees/crevasse-splay, slumps, channels (Upper Anguille); Table 2
gives mean petrophysical characteristics. A very important final result for reservoir
simulation is the identification of five scales of heterogeneity - Figures 6 and 7; this
makes the geological analysis and information numerically useable.
7. Sedimentary history: In earlier reservoir simulation studies, and indeed up to the
present time, it is rare to see sedimentary history discussed in terms of a sequence
stratigraphic analysis (even mentioning the pioneering work on sea level changes of
P. R. Vail et al, Seismic stratigraphy and global changes of sea level, in Seismic
Stratigraphy, Applications to Hydrocarbon Exploration, AAPG Memoir 26, pp. 49-212,
1977). Chronostratigraphic correlations refer to the timelines of simultaneous
deposition. This analysis underpins much of the reservoir description but we will
not elaborate on it here.
8. Geostatistical modelling: Mainly discussed in Refs. 1 and 2 of this paper. Firstly,
focus on geostatistical modelling of the 3D distributions of the major flow units
(channels and lobes) and barriers (laminated facies or slumps) for the entire reservoir.
This is done as a conditional simulation where the distribution is constrained
30

Introduction and Case Studies

(or conditioned) to the observed facies and reservoir quality observed at the
wells. Secondly, the smaller scale heterogeneities are unconditionally simulated
(synthetically) to yield average properties within the major flow units (see Ref. 2 in
paper). The geostatistical simulation method used was indicator simulation (Refs.
1 and 8) which require the average frequency and variogram information.
9. Reservoir zonation: Six unit vertical reservoir zonation shown in Figure 10.
Simulations of lateral continuity within each of the units (five - not Middle Anguille,
Figure 10) performed independently since they correspond to separate sedimentary
phases. For horizontal zonation, Figure11 shows lateral zonation on LA2
and UA2 units showing directional trends and thus variograms with spatially
variable anisotropy direction used in final model. Figures 12 - 15 show resulting
correlation structures of the various units. Ends up with >2 million grid blocks
in the full field 3D model.
10. Flow simulations: Discusses details of upscaling from fine grid stochastic model
(>2 million blocks) to coarse grid simulation model (11,000 grid blocks). 11 vertical
layers are retained to represent the reservoir layering with more blocks being used
in the best reservoir units. Upscaling of absolute permeability at some aggregation
rate (e.g. 4x4) is applied leading to areal block sizes of 200m x 200m - see Figure14.
Relative permeabilitiees were upscaled on a typical block configuration (details
in Refs. 2 and 4). Additionally: Three major zero-transmissibility faults included in
model; some WOC variation across field; depth varying bubble points assigned; 25
years of injection/production for history matching.
11. Simulation results: Initial pressure depletion results shown in Figure 16 - where
14 out of 17 wells show satisfactory pressure behaviour. Pressure behaviour and
water breakthrough are poorly predicted during injection stage - Figure 17; water
saturations around injectors shown in Figure18 - upscaling has washed out the
finer scale strong anisotropy.
12. Model changes: Table 8 lists a number of sometimes quite radical changes to
the model in order to achieve a better fit to observed field performance - Figure 15
shows differences in upscaled permeability maps. Continuing problems with injection
predictions => - is geological model correct? - what is the real effect of upscaling?
13. Partial models: Thin model - Figure 19 shows the thin partial field model to
verify reservoir geology; well AGM18 good water breakthrough match (Figure 20)
- early breakthrough for well AGM29 (Figure 21). When thin model upscaled as in
full field model (abs. k upscale + rel perm as before) - results in Figures 21 and 22
- breakthrough delayed in both wells but shape of BSW is satisfactory. Conclusions:
Thin model partly validates geological model; Some problems with upscaling not
supressing breakthrough, making reservoir too connected and eliminating strong
anisotropy. Test model - (50 x 20 x 56) model extracted from full field model. Figure
23 shows that an optimum upscaling aggregation rate (2 x 2 x 7) is found - they warn
caution on this point. We note that if very reliable and general upscaling techniques
were available, then this should be eliminated (more work has been done on this
issue since 1992 - much of it at Heriot-Watt!).

Institute of Petroleum Engineering, Heriot-Watt University

31

14. Conclusions:
Sedimentology controls heterogeneity analysis when very wide variation in
sandbody geometries is found (as in this case)

Link understanding of reservoir history to sequene stratigraphy

Litho-interpretation of seismic cant give paleo-direction when there is techtonic


activity during sedimentation

Multi-scale heterogeneity analysis essential to quantify sub-grid petrophysical


properties

Geostatistical indicator simulation is a good tool for modelling this multi- scale
heterogeneity - trends can also be included

Stochastic model for Anguille Marine constrained by geology gives hopeful


first results

If aggregation rate in upscaling is optimised, history matching is possible with


the use of strictly controlled geological parameters

3.6 Case 3: Ubit Field Rejuvenation (SPE49165,1998)

Case 2: The Ubit Field Rejuvenation: A Case History of Reservoir Management


of a Giant Oilfield Offshore Nigeria, SPE49165, presented at the SPE Annual
Technical Conference and Exhibition, New Orleans, LA, 27-30 September 1998,
by C.A. Clayton et al (Mobil and Department of Petroleum Resources, Nigeria)
See SPE 49165 paper in Appendix
Summary: This is another good example of where integrated reservoir management
has greatly contributed to the success of a field redevelopment plan. In particular,
a clearer understanding of the reservoir structural geology has been central to this
process. The reinterpretation of the structural geology of the field (the fault blocks,
compartments and slump blocks) was achieved using seismic data in a range of
complementary ways. The Ubit reservoir is a prograding shallow marine system
which has been tectonically disturbed. The downslope movements of the youngest
sand sequences resulted in large scale slumping and block sliding although reservoir
quality in these sediments is good to excellent. Important facts on the Ubit reservoir
and this study are: STOIIP = 2.1 billion bbl oil; 37API black oil, Bo = 1.38, GOR 612
scf/stb, o = 0.64 cp and g = 0.16 cp; production from a relatively thin oil column
(160 ft.) and a fairly thick gas cap (50 - 550 ft.). Previous average production = 30
MBD; after implementation of study recommendations (many horizontal wells etc.),
expected production 140 MBD.
The notes on this SPE paper will not be very extensive and only a few of the main
novel points will be discussed below.

Study Notes Case 3:

1. New data and techniques: The study is a very good example of the close
integration of (especially) 3D seismic data used in several ways, computer mapping
32

Introduction and Case Studies

and reconstruction of the slump blocks, advanced reservoir simulation procedures,


visualisation etc.
2. Recommendations: These have been quite clearly established and stated. The
study shows that the key strategies are

Implement horizontal well drilling (approx. 57 wells).


Full field simulation defining drilling placement and timing.
Balancing a non-uniform gas cap.
Maintaining a stable gas cap (gravity stable displacement) and pressure.
Establish field plateau rate.
Minimising free gas production.

3. Uncertainties: there was an initially erroneous view of certain aspects of the


reservoir geology and the key uncertainties at the start of the study were

Geological complexities in reservoir architecture, particularly structural


deformation.
Sandbody geometries.
Petrophysical rock and fluid properties.
Distribution of flow units.

4. Structural reinterpretation: Figures 2a and 2b show both the original and current
interpretations of the structure. The original rubble beds are reinterpreted as being
techtonically disturbed downslope movements of the youngest sand sequences
resulting in large scale slumping and block sliding. The older interpretation saw
these facies as being essentially chaotic whereas they are now thought to be more
ordered and predictable. 3D seismic data is of central importance in the definition
of the structural geometries where the bedded and disturbed strata are shown
on a seismic section in Figure 4. Several seismic techniques were applied including
attribute analysis, rock physics and amplitude analysis, seismic facies analysis of
time slices and conventional reflector mapping. The resulting 70 internal slump
and fault blocks are shown in Figure 7.
5. Petrophysics-based facies: Seven flow controlling depositional facies were
identified as shown in Figure 8 with rock properties related to grain size (lithology,
typical log response, net/gross, k vs. , Pc and kro-krw). Depositional facies types
present are - marine turbidites and debris flow sands; lower delta plain tidal channels
and lagoonal sands; shallow marine upper shoreface and lower shoreface sands and
shelf shales (best are turbidites, shoreface and channel sands - comprise 80% pore
volume in oil column).
6. Layering and Reservoir Simulation Model: Vertical layering is shown
schematically in Figure 9 and the areal grid is given in Figure 12, with a set of rock
property maps for a single simulation layer given in Figure10. Grid is
Nx x Ny x Nz = 93 x 40 x 18 (67,000 blocks) with most oil leg cells being z = 10ft.
to resolve the gravity stable gas front. Rock property slices were loaded into the 3D
modelling software to connect up the stratigraphic layers (using new but unclear
developments by authors) as shown in Figure 11.

Institute of Petroleum Engineering, Heriot-Watt University

33

7. Relative permeability: An interesting point is made concerning the oil relative


permeability (kro) close to its end point - see Figure 14. Although kro is low in this
region (kro ~10-6 - 10-7), it significantly affects the tail of the reservoir oil production
profile - see Figure 15. The adjusted curve in Figure 14 is used to get more realistic
longer time recoveries (Figure 15). This is important because gravity stable gas cap
expansion and downward displacement is the principal oil recovery mechanism - see
schematic view in Figure 9.
8. History match and forward prediction: Average field pressure and GOR are
history matched - see Figures 15 and 16. Matching field pressure is not so difficult
since Ubit shows good pressure communication (high k - lack of sealing faults).
Some discussion of field management is presented. Figure 22 shows the initial part
of the improved productivity (up to 140 MBD from 37 horizontal wells) and future
predictions. See Recommendations (point 2 ) above.

3.7 Discussion of Changes in Reservoir Simulation; 1970s - 2000

From the above field examples (Cases 1 -3), there is clearly a progression in the
engineering approach, the degree of reservoir description and the computational
capabilities as we go from reservoir simulation in the late 1970s to the present
time.
The main changes are as follows:
Computer power: There has been a vast increase in computer processing power
over this period because of :
(a) CPU: The growth of powerful CPU (central procesing units - i.e. chips) especially
as implemented in Unix machines (workstations) and RISC technology and
more recently by the development of modern PCs. The corresponding cost of
computing has fallen dramatically. A graph of processing power (Mflops/s) vs.
time and a corresponding graph of maximum practical model size vs. time is
shown in Figure 10:

1000000

Gridblocks

Mflops/s

1000

100000

100
10
1

1000

0.1
1970

10000

1975

1980

1985

1990

1995

Year
(a) State of art CPU performance

2000

100
1960

1970

1980

Year

1990

2000

(b) Maximum practical simulation model size

(b) Parallel Processing: Part of the increase in computing power referred to above
is the growth of parallel processing in reservoir simulation. The central idea
here is to distribute the simulation calculation around a number of processors
( or nodes) which perform different parts of the computational problem
simultaneously. A bank of such processors is shown in Figure 11 (from the
34

Figure 10
(a) CPU performance
(Mflops/s) vs. time and (b)
maximum practical model
size vs. time; Mflop/s =
mega-flops per second =
million floating point
operations per second; from
J.W. Watts, Reservoir
Simulation: Past, Present
and Future, SPE Reservoir
Simulation Symposium,
Dallas, TX, 5-7 June 1997.

Introduction and Case Studies

work of Dogru, SPE57907, 2000). The general impact of parallel simulation


is shown according to Dogru (2000) in Figure 12. If the problem gets linearly
faster with the number of parallel processors, then it is said to be scalable and
the closeness to an ideal line is a measure of how well the process parallelises
(reaches the ideal scaling line); an example is shown in Figure 13. Finally,
the type of fine scale calculation that can now be performed using megacell
simulation is shown in Figure 14 where it is shown that there is a lengthscale
of remaining oil that is missed in the coarser (but still quite fine) simulation.
A table of what types of calculation can be performed and some timings for
these is also included (although these numbers will probably be out of date
very quickly!). For further details, see Megacell Reservoir Simulation - A.H.
Dogru - SPE Distinguished Author Series, SPE57907, 2000 and the references
therein.

Figure 11
A cluster of parallel
processors; from Megacell
Reservoir Simulation A.H. Dogru - SPE
Distinguished Author
Series, SPE57907, 2000.

1.2
1

Parallel Simulators

0.8

Figure 12
The impact of parallel
reservoir simulation; from
Megacell Reservoir
Simulation - A.H. Dogru
- SPE Distinguished Author
Series, SPE57907, 2000.

0.6
0.6
Conventional Simulators

0.2
0

1988

1990

1992

1994

1996

Institute of Petroleum Engineering, Heriot-Watt University

1997

1998

35

0.6
Conventional Simulators

0.2
0

1988

1990

1992

1994

1996

1997

1998

Speedup

16

12

12

Figure 13
A cluster of parallel
processors; from Megacell
Reservoir Simulation A.H. Dogru - SPE

16

Number of Nodes
Cluster

SP2

Ideal

Distinguished Author Series,


SPE57907, 2000.

Fig. 7Comparison of conventional simulation (40,500


cells, 5 la
yers) and megacell simulation 2.45
(
million
cells,67 layers).

Reservoir

Model Size
(Millions of Gridblocks)

History
Length (Years)

CPU Hours
On 64 Nodes On 128 Nodes

Carbonate

1.2

27

1.7

Sandstone

1.3

49

4.5

2.5

Carbonate
(With gas cap)

3.9

10

2.0

Carbonate

2.5

2.5

4.0

(c) Visulisation: Huge improvements in visualisation capabilities have taken which


allow us to evaluate vast quantities of numerical data in a more convenient
manner. This, in turn allows us to apply our intuitive engineering judgement
to reservoir development problems. In addition, the visual representation of
output allows a more fruitful communication to take place between geologists
and engineers;
(d) Integrated software: The availability of more integrated software suites that
handle all the initial data from seismic processing, to petrophysical analysis
and data generation, geocellular modelling and upscaling through to the actual
simulators themselves.
36

Figure 14
The type of reservoir
simulation that becomes
more possible with parallel
processing. Comparison
with fine grid and megacell
simulation which identifies
the scale of remaining oil
in a reservoir displacement
process; from Megacell
Reservoir Simulation A.H. Dogru - SPE
Distinguished Author
Series, SPE57907, 2000.

Introduction and Case Studies

(e) Linked to this increased power, is the ability to handle huge geocellular models
and somewhat smaller but still very large reservoir simulation models.
Geostatistics: there have been significant advances in the application of geostatistical
techniques in reservoir modelling. Such approaches were quite well known in the
mining, mineral processing and prospecting industries but only in the last 10 to 15
years have they been specifically adapted for application in petroleum reservoir
modelling. Introductory texts are now available such as:

An Introduction to Applied Geostatistics by E.H. Isaaks and R.M. Srivastava,


Oxford University Press, 1989

Introduction to Geostatistics: Applications in Hydrogeology by P.K. Kitanidis,


Cambridge University Press, 1997

Both pixed-based point geostatistical techniques and object based modelling have
been developed and applied in various reservoirs.
Upscaling: There have been a number of advances in approaches to upscaling (or
pseudo-isation) from fine geocellular model the reservoir simulation model. This
still an area of active research and the debate is still in progress on the question:

Will increasing computing power remove the need for upscaling?

The basic idea of upscaling has been introduced in the SPE examples. Upscaling is
dealt with in much more detail in Chapter 7.
Organisational changes in the oil industry: A number of major organisational
changes have occurred in the oil industry since the 1970s which have affected the
practice of reservoir simulation. The main ones are as follows:
(a) Many companies have taken a more integrated geophysics/geology/engineering
view of reservoir development and many studies have made a central virtue of this
by organising reservoir studies within more multi-disciplinary asset teams (e.g.
SPE25006 clearly shows a strong integration of geology and engineering);
(b) There have been significant organisational changes in the structure of the
industry given the sucessive rounds of downsizing and outsourcing that have
occurred. For example, see the short article by Galas, The future of Reservoir
Simulation, JCPT, p.23, Vol. 36 (1), January 1997, which is reproduced in
Appendix B. This article has an interesting slant from the point of view of the
smaller consultant and it makes a number of interesting observations;
(c) There have been a number of major mergers and take-overs recently which
have formed some very large companies e.g. BP (BP - Amoco - ARCO), ExxonMobil, Total-Fina-Elf. Likewise, a number of very low cost operations have
grown up which may specialise in the successful (i.e. profitable) exploitation of
mature assets e.g. Talisman, Kerr-McGee etc. How these changes will affect
the future of reservoir simulation remains to be seen.

Institute of Petroleum Engineering, Heriot-Watt University

37

(d) Up until the late 1970s, almost every major and medium sized oil company
had a research centre where programmes of applied R&D were carried
out by oil company personnel. The view was essentially that this in-house
technology development would give the company a competitive and commercial
edge in reservoir exploration and development. Most companies have greatly
reduced the amount of in-house R that takes place and have focused much
more heavily on shorter term asset related D. Many companies do support
research in universities and other independent outside organisations - they also
ally themselves with service companies in order to have their R&D needs met
in certain areas. Again, the situation is in flux and the longer term effects of
this change is yet to be seen.
(e) More specific to reservoir simulation is the fact that, in the 1970s, most
companies would have had their own numerical reservoir simulator which was
built (programmed up) and maintained in-house. To this day, a few companies
still do. However, most oil companies use specialised software service companies
to supply their reservoir simulation (and visualisation, gridding etc.) software.
Again, the relative merits and demerits of this will emerge in the coming
years.
Detailed technical advances: In addition to the changes discussed above, many
advances have been made over the past 50 years on how we perform the simulations
i.e. on the formulation and numerical methods etc. Our practical capabilities have
also expanded greatly as discussed above. Table 1 presents a list of capabilities and
major technical advances in reservoir simulation over the last 50 years; this table
was adapted from two tables in J.W. Watts Reservoir Simulation: Past, Present and
Future, SPE38441, SPE Reservoir Simulation Symposium, Dallas, TX, 5-7 June
1997. This article is well worth reading. Most of the technical details in the advances
listed in Table 1 are beyond the scope of this course and the introductory student does
not need to have any in-depth knowledge on these.

38

Introduction and Case Studies

Decade

Capabilities

References

1950s

Two dimensions
Two incompressible phases
Simple Geometry

Aronofsky and Jenkins radial gas model


Alternating-direction implicit (ADI) procedure

Aronofsky, and Jenkins (1954)


Peaceman and Rachford (1955)

1960s

Three dimensions
Three phases
Black-oil fluid model
Multiple wells
Realistic geometry
Well coning

IMPES computational method


Upstream weighting
Understanding of numerical dispersion
Strongly-implicit procedure (SIP)
Implicit computational method
Additive correction to line-successive
overrelaxation

Sheldon et al (1960)
Stone and Garder (1961)
Lantz (1971)
Stone (1968)

Compositional
Miscible
Chemical
Thermal

Stone relative permeability models


Vertical equilibrium concept
Todd-Longstaff miscible displacement
computation
Two-point upstream weighting
D4 direct solution method
Total velocity sequential implicit method
Pseudofunctions
Variable bubblepoint black-oil treatment
Conjugate gradients and ORTHOMIN

Stone (1970, 1973)


Coats et al (1971)
Todd and Longstaff (1972)

1970s

Table 1
Capabilities and major
technical advances in
reservoir simulation over
the last 50 years (adapted
from two tables in J.W.
Watts Reservoir
Simulation: Past, Present
and Future, SPE38441,
SPE Reservoir Simulation
Symposium, Dallas, TX, 5-7
June 1997; (all references
are given in Appendix A)

Technical Advances

Iterative methods based on approximate


factorizations
Peaceman well correction
Nine-point method for grid orientation effect

MacDonald and Coats (1970)


Watts (1971)

Todd et al (1972)
Price and Coats (1974)
Spillette et al (1973)
Kyte and Berry (1975)
Thomas et al (1976)
Meijerink and Van der Vorst
(1977)
Vinsome (1976)
Peaceman (1978)
Yanosik and McCracken (1979)

1980s

Complex well management


Fractured reservoirs
Special gridding at faults
Graphical user interfaces

Code vectorization
Nested factorization
Volume balance formulation
Young-Stephenson formulation
Adaptive implicit method
Constrained residuals
Local grid refinement
Cornerpoint geometry
Geostatistics
Domain decomposition

Appleyard and Cheshire (1983)


Acs et al (1985); Watts (1986)
Young and Stephenson (1983)
Thomas and Thurnau (1983)
Wallis et al (1985)
Ponting (1992)

1990s

Improved ease of use


Geologic models and upscaling
Local grid refinement
Complex geometry
Integration with non-reservoir
computations

Code parallelization
Upscaling
Voronoi grid

Heinemann et al (1991)
Palagi and Aziz (1994)

3.8 The Treatment of Uncertainty in Reservoir Simulation

Figure 15
A single forward prediction
of the oil recovery
production profile for a
given reservoir.

Cumulative Oil Recovery

It is well recognised in modern reservoir development that when we calculate the


future oil recovery profile of a reservoir, it is not accurate. Suppose a particular
reservoir simulation model for Field X - a new development which is currently under
appraisal - gives the forecast in Figure 15.

2000

2005

2010
Time (Year)

Institute of Petroleum Engineering, Heriot-Watt University

39

Clearly, we cannot trust this single curve since there is a considerable amount of
uncertainty associated with it for various easily appreciated reasons. The main
contributors to this uncertainty are to do with lack of knowledge about the input
data although the modelling process itself is not error free. A list of possible sources
of error is as follows:

Lack of knowledge or wide inaccuracies in the size of the reservoir; its areal
extent, thickness and net-to-gross ratios

Lack of knowledge about the reservoir architecture i.e. its geological structure
in terms of sandbodies, shales, faults, etc.

Uncertainties in the actual numerical values of the porosities () and permeabilities


(k) in the inter-well regions (which make up the vast majority of the reservoir
volume)

Inaccuracy in the fluid properties such as viscosity of the oil (o), formation volume
factors (Bo, Bw, Bg), phase behaviour etc., or doubts about the representativity
of these properties

Lack of data - or very uncertain data- on the multiphase fluid/rock properties,


particularly relative permeability and capillary pressure, and on knowledge as
to how these curves vary from rock type within the reservoir volume away from
the wells

Because the representational reservoir simulations model may be poor, e.g. the
numerical errors due to the coarse grid block model may significantly affect the
answer in either an optimistic or pessimistic manner.

The above list of uncertainties for a given reservoir, especially at the appraisal stage,
is really quite realistic and is by no means complete. As we have noted elsewhere,
it is at the appraisal stage when, although the future reservoir performance is at its
most uncertain, we must make the biggest decisions about the development and hence
speed most of our investment money.
At a first glance, the task of doing something useful with reservoir simulation may
seem quite hopeless in the face of such a long list of uncertainties. No matter how
bleak things look, the only two options are to give up or do something, and reservoir
engineers never give up! We must produce an answer - even if it is an educated
guess (or even just a guess) - and some estimate of the sort of error sound that we
might expect.
Before considering what we can do in practice, let us first consider what the answer
might look like for the case above in Figure 15. Figure 16 gives some idea of what
is required:

40

Introduction and Case Studies

Figure 16
Outcome of reservoir
simulation calculations
showing a range of
recoveries for various
reservoir development
scenarios.

Cumulative Oil Recovery

Most probable case

Range of cases with


a 50% probability

Range of cases with


a 90% probability

2000

2005

2010
Time (Year)

The results in Figure 16 can be understood qualitatively without worrying about how
we actually obtain them right now. Our single curve in Figure 15 may becomes a
most probable (or base case) future oil recovery forecast. The closer set of outer
curves is the range of future outcomes that can be expected with a 50% probability.
That is, there is a probability, p = 0.5, that the true curve lives within this envelope
of curves shown in Figure 16. Such results allow economic forecasts to be made with
the appropriate weights being given to the likelihood of that particular outcome. A
company can then estimate its risk when it is considering various field development
options.
In fact, here we will just discuss doing some simple sensitivities to various factors in
the simulation model. We can think of a given calculation as a scenario. Therefore,
we can set up various scenarios based on our beliefs about the various input values
in our model and we simply compare the recovery curves for each of the cases. For
example, suppose we have a layered reservoir as shown in Figure 17 which we think
has a field-wide high permeability streak set in background of 100 mD rock.
INJECTOR

Figure 17
This shows a layered
reservoir where we have
some uncertainties in the
various parameters such as
the permeability (khi ), the
thickness (Zhi ) and the
porosity (hi ) in the high
permeability layer.

PRODUCER
1000ft
1000ft

klow = 100mD

100ft
= 0.18
High Permeability Streak,
khigh, hi

Zhi

The reservoir is being developed by a five-spot waterflood as shown in Figure 17


However, we are uncertain as to the actual thickness (Zhi), the permeability (khi) and
the porosity (hi) of the high permeability streak. From various sources, we derive
mean, high and low estimates of each of these quantities as follows:

Institute of Petroleum Engineering, Heriot-Watt University

41

Permeability, khi
Thickness, Zhi
Porosity, hi

Low Value

Mean Value

High Value

400 mD

800 mD

1600 mD

20 ft
0.18

30 ft
0.22

40 ft
0.26

Even with just the three uncertainties in this single model, we can see that there are
3x3x3 = 27 possible scenarios or combinations of input data for which we could
run a reservoir simulation model. Alternatively, we could conclude that some input
combinations are unlikely (e.g. lower permeability with higher value of porosity) and
we could reduce the number. We could simply keep the mean value of two of the
factors while varying only the third factor, leading to 7 scenarios to simulate. Taking
this view, we can take some measure of the oil recovery e.g. cumulative oil produced
(predicted) at year 2010. The notional results could be entered in Table 2
Changed Input
Value

Oil Initially in

Cumulative

% Change in

Change in Recovery

Place (OIIP)
(res. bbl)

Recovery at
Year 2010 (stb)

Input Value Relative


to Base Case

from Base Case

Base Case
khi = 400 mD
khi = 1600 mD

Table 2
Results of sensitivity
simulations described in the
text.

Zhi = 20 ft
Zhi = 40 ft
hi = 0.18
hi = 0.25

Note: The OIIP will vary somewhat from case to case since the thickness of the high
permeability layer and its porosity both change.
In Table 2, we have noted the % change in the varied parameters relative to its base
case value. Not that different physical quantities such as k and , vary by different
percentages for realistic min./max. values. A useful way to plot the variation in recovery
is against this % change in input value since all three factors can be represented on the
same scale in a so-called spider diagram. Such a plot is shown in Figure 18.

Porosity

Permeability

X
X

Layer Thickness

% Change in Parameter

Change in Recovery From Base Case (STB)

42

Figure 18
Spider diagram showing
the sensitivity of the
cumulative oil to various
uncertainties in the
reservoir model parameters
(khi; Zhi; hi) in Figure 17.

Introduction and Case Studies

This type of spider plot is very useful since it displays the effect of the different
uncertainties on the outcome. It clearly highlights which is the most important input
quantity (of those considered) and has the most impact on the result. Thus, if we
were going to spend time and effort on reducing the uncertainty in our predictions,
then this tells us which quantity to focus on first. Indeed it ranks the effects of the
various uncertainties.
There are more sophisticated ways to deal with uncertainty in reservoir performance
but these are beyond the scope of the current course. The basic ideas presented above
give you enough to go on with in this course.

4 STUDY EXAMPLE OF A RESERVOIR SIMULATION


The examples presented in the above SPE papers should give you a good idea of
what reservoir simulation is all about. By this point you should also be familiar
with many of the basic concepts that are involved in reservoir simulation from the
study notes and the Glossary. However, the reservoir simulation of say a waterflood
or a gas displacement is a dynamic process. That is, as time progresses, the water
or gas front should be moving through the reservoir interacting with the underlying
geological structure in quite a complicated manner. The pressure distribution through
the system should also be evolving with time. For example, the water may possibly
be advancing preferentially through a high permeability channel between an injector
and producer pair. The sequence of the saturation fronts in the series of snapshots
in time shown in Figure 19 give some idea of the progression of the flood with time.
However, this can better be illustrated by looking at an animated sequence of saturation
distributions as the flood front moves through the reservoir. An example of such a
sequence is shown in file Res_Sim_D1.ppt This can be run on your PC from the
CD supplied with this course and double clicking on the PowerPoint presentation.

Institute of Petroleum Engineering, Heriot-Watt University

43

Figure 19
The sequence of saturation
distributions as the flood
front moves through the
reservoir. From Res_
Sim_D1.ppt Down arrow
injector, up arrow producer.

44

Introduction and Case Studies

5 TYPES OF RESERVOIR SIMULATION MODEL


Until now, we have confined our discussion to relative simple reservoir recovery
processes such as natural depletion (blowdown) and waterflooding. However, there
are many more complex reservoir recovery processes which can also be carried out.
Dry gas (methane, CH4) injection, for example, would generally result in the flow
of three phases (gas, oil and water) in the reservoir which is more complicated than
two phase flow. Another process is where we alternately inject water and gas in
repeating sequence - this is water-alternating-gas or WAG flooding. If the injected
gas was carbon dioxide (CO2), then quite complex phase behaviour may occur and
this requires some particular steps to be taken in order to model this. More exotic
Improved Oil Recovery (IOR) processes can also be carried out where we inject
chemicals (polymers, surfactants, alkali or foams) into the reservoir to recover oil
that is left behind by primary and secondary oil recovery processes.

5.1 The Black Oil Model

Different types of simulator are available to model these different types of reservoir
recovery process. Throughout the chapters of this course we will focus on the
simplest of these (which is quite complex enough!) known as the "Black Oil Model".
However, for completeness, we will also list the others and present a table comparing
experience of these various models.
The Black Oil Model: This model was used in the three SPE field case studies above
and is the most commonly used formulation of the reservoir simulation equations
which is used for single, two and three phase reservoir processes. It treats the three
phases - oil, gas and water - as if they were mass components where only the gas
is allowed to dissolve in the oil and water. This gas solubility is described in oil
and water by the gas solubility factors (or solution gas-oil ratios), Rso and Rsw,
respectively; typical field units of Rso and Rsw are SCF/STB. These quantities are
pressure dependent and this is incorporated into the black oil model.
A simple schematic of a grid block in a black oil simulator is presented in Figure
20 showing the amounts of mass of oil, water and gas present. Note that, because
the gas is present in the oil and water there are extra terms in the expression for the
mass of gas. These mathematical expressions for the mass of the various phases are
important when we come to deriving the flow equations (Chapter 5).
Reservoir processes that can be modelled using the black oil model include:

Recovery by fluid expansion - solution gas drive (primary depletion).

Waterflooding including viscous, capillary and gravity forces (secondary


recovery).

Immiscible gas injection.

Some three phase recovery processes such as immiscible water-alternatinggas (WAG).

Capillary imbibition processes.

Institute of Petroleum Engineering, Heriot-Watt University

45

Vp.osc
Mass oil =
Bo

So;

Mass water =

Mass gas =

Vp.wsc
Bw

Sw

Vp.gsc
(Sg + So.Rso + Sw.Rsw)
Bg

Rock
Gas, Sg
Oil + Gas, So
Water + Gas, Sw

free gas
gas in oil
gas in water

5.2 More Complex Reservoir Simulation Models

The Compositional Model: A compositional reservoir simulation model is required


when significant inter-phase mass transfer effects occur in the fluid displacement
process. It can be considered as a generalisation of the black oil model. This model
usually defines three phases (again gas, oil and water) but the actual compositions
of the oil and gas phases are explicitly acknowledged due to their more complicated
PVT behaviour. That is, the separate components (C1, C2, C3, etc.) in the oil and gas
phases are explicitly tracked as is indicated in Figure 21 (which should be compared
with Figure 20). The mass conservation is applied to each component rather than
just to oil, gas and water as in the black oil model. For example, in a nearcritical fluid where small changes in say pressure can result in large compositional
changes of the oil and gas phases which, in turn, strongly affects their physical
properties (viscosity, density, interfacial tensions etc.).
Examples of reservoir processes that can be modelled using a compositional model
include:

Gas injection with oil mobilisation by first contact or developed (multi- contact)
miscibility (e.g. in CO2 flooding).

The modelling of gas injection into near critical reservoirs.

Gas recycling processes in condensate reservoirs.

46

Figure 20
Schematic of a grid block
in a black oil simulator
showing the amounts of
mass of oil, water and gas
present. Note that, because
the gas is present in the oil
and water there are extra
terms for the mass of gas;
pore volume = Vp = block
vol. x ; osc, wsc. and gsc
are densities at standard
conditions (60F and 14.7
psi); Bo, Bw and Bg are the
formation volume factors;
Rso and Rsw are the gas
solubilities (or solution
gas/oil ratios).

Introduction and Case Studies

Component concentrations
in each phase:

ROCK

Figure 21
The view of phases and
components taken in
compositional simulation.
Cij - is the mass
concentration of component
i in phase j (j = gas, oil or
water) - dimensions of mass/
unit volume of phase; pore
volume = Vp = block vol. x

Phase Labels:
j = 1 = Gas
j = 2 = Oil
j = 3 = Water

GAS
Sg
OIL + GAS
So
WATER + GAS
Sw

Gas:

C11, C21, C31....

Oil:

C12, C22, C32....

Water: C13, C23, C33....

Mass of component in block = Vp. Sj.Cij


j=1

The Chemical Flood Model: This model has been developed primarily to model
polymer and surfactant (or combined) displacement processes. Polymer flooding
can be considered mainly as extended waterflooding with some additional effects in
the aqueous phase which must be modelling e.g. polymer component transport, the
viscosification of the aqueous phase, polymer adsorption, permeability reduction
etc. Surfactant, flooding however, involves strong phase behaviour effects where
third phases may appear which contain oil/water/surfactant emulsions. Specialised
phase packages have been developed to model such processes. For economic reasons,
activity on field polymer flooding has continued at a fairly low level world wide
and surfactant flooding has virtually ceased in recent years. However, if economic
factors were favourable (a very high oil price), then interest in these processes may
revive. Extended chemical flood models are also used to model foam flooding.
Examples of reservoir processes that can be modelled using a chemical flood model
include:

Polymer flooding which can be thought of as an enhanced waterflood to


improve the mobility ratio and hence improve the microscopic sweep efficiency
and also to reduce streaking in highly heterogeneous layered systems;

Polymer/surfactant flooding where the main purpose of the surfactant is to


lower interfacial tension (IFT) between the oil and water phases and hence to
release or mobilise trapped residual oil; the polymer is for mobility control
behind the surfactant slug;

Low-tension polymer flooding (LTPF) where a more viscous polymer containing


injected solution also contains some surfactant to reduce IFT; the combined
effect of the lower IFT and viscous drive fluid improves the sweep and also
helps to mobilise some of the residual oil;

Alkali flooding where a solution of sodium hydroxide is injected into the


formation. The sodium hydroxide may react with certain conponents in the oil
to produce natural "soaps" which lower IFT and which may help to mobilise
some of the residual oil;

Institute of Petroleum Engineering, Heriot-Watt University

47

Foam flooding where a surfactant is added during gas injection to form a foam
which has a high effective viscosity (lower mobility) in the formation than the
gas alone which may then displace oil more efficiently.

Another near-wellbore process that can be modelled using such simulators in water
shut-off using either polymer-crosslinked gels or so-called relative permeabilty
modifiers.
Thermal Models: In all thermal models heat is added to the reservoir either by
injecting steam or by actually combusting the oil (by air injection, for example). The
purpose of this is generally to reduce the viscosity of a heavy oil which may have o
of order 100s or 1000s of cP. The heat may be supplied to the reservoir by injected
steam produced using a steam generator on the surface or downhole. Alternatively,
an actual combustion process may be initiated in the reservoir - in-situ combustion
- where part of the oil is burned to produce heat and combustion gases that help to
drive the (unburned) oil from the system.
Examples of reservoir processes that can be modelled using thermal models
include:

Steam soaks where steam in injected into the formation, the well is shut in for
a time to allow heat dissipation into the oil and then the well is back produced
to obtain the mobilised oil (because of lower viscosity). This is known as a
Huff n Puff process.

Steam drive where the steam is injected continuously into the formation
from an injector to the producer. Again, the objective is to lower oil viscosity
by the penetration of the heat front deep into the reservoir.

In situ combustion where - as noted above - an actual combustion process is


initiated in the reservoir by injecting oxygen or air. Part of the oil is burned
(oxidised) to produce heat and combustion gases that help to drive the (unburned)
oil from the system. This is not a common improved oil recovery method but a
number of field cases showing at least technical success have been reported in
the SPE literature.

The above more complex reservoir simulation models are really based on the fluid flow
process. However, there are also other types of simulator that are more closely defined
by their treatment of the rock structure or the rock response. These include:
Dual-Porosity Models of Fractured Systems: These models have been designed
explicitly to simulate multiphase flow in fractured systems where the oil mainly
flows in fractures but is stored mainly in the rock matrix. Such models attempt to
model the fracture flows (and sometimes the matrix flows) and the exchange of fluids
between the fractures and the rock matrix. They have been applied to model recovery
processes in massively fractured carbonate reservoir such as those found in many
parts of the Middle East and elsewhere in the world. There is quite considerable field
experience of modelling such systems in certain companies but there are also doubts
over the validity of such models to model flow in fractured systems.
48

Introduction and Case Studies

Coupled Hydraulic, Thermal Fracturing and Fluid Flow Models: These simulators
are still essentially at the research stage although there have been published examples
of specific field applications. The main function of these is to model the mechanical
stresses and resulting deformations and the effects of these on fluid flow. This is
beyond the scope of this course although, in the future, these will be important in
many systems.

5.3 Comparison of Field Experience with Various Simulation Models

We now consider what field experience exists in the oil industry with the various
models from the black oil model through to more complex fracture models and in situ
combustion models etc. The vast majority of simulation studies which are carried
out involve the black oil model. However, there are pockets of expertise with the
various other types of simulation model, depending on the asset base of the particular
oil company or regional expertise within regional consultancy groups. For example,
there is (or until recently, was) a concentration of expertise in both California and
parts of Canada on steam flooding since this process is applied in these regions to
recover heavy oil; in the Middle East (and within the companies that operate there)
there is great competence in the dual-porosity simulation of fractured carbonate
reservoirs.

Table 3
This is an adapted version
of a table in Chapter 11 of
Mattax and Dalton (1990).
This gives some idea of
the problems and issues
encountered in applying
advanced simulation models
relative to applying a black
oil simulator. The view
about the difficulties and
computer time consuming
these are is somewhat
subjective.

Degree of Difficulty

Relative Computing
Costs

Processes Modelled

Black Oil Model

Primary depletion
Waterflooding
Immiscible gas
injection
Imbibition

Routine

Cheap = 1

Huge
But there are still
challenges with
upscaling of large
models
>90% of cases

Any of the books on


reservoir
simulation listed in
Section 7 (Chapter 1)

Compositional
Model

Gas injection
Gas recycling
CO2 injection
WAG

Difficult
Specialisd

Expensive
(x3 - x20)

Moderate
High in certain
companies

Coats, (1980a),
Acs et al (1985),
Nolen (1973),
Watts (1986),
Young and Stephenson
(1983).

Compositional
Model- Near Crit.

Gas injection near


crit.
Condensate
development
MWAG

Difficult

Very expensive
(x5 - x30)

Low to moderate

Chemical Model
- Polymer

Polymer flooding
Near-well water
shut-off

Not too difficult

Moderate
(x2 - x5)

Moderate to large

Bondor et al (1972),
Vela et al (1976),
Sorbie (1991)

Chemical Model
- Surfactant

Micellar flooding
Low tension polymer
flooding

Difficult
Specialisd

Expensive
(x5 - x20)

Low
Mainly research type
pilot floods

Todd and Chase (1979),


Todd et al (1978),
Van Quy and Labrid (1983);
Pope and Nelson (1978)

Thermal Model
- Steam

Steam soak
(Huff n Puff)
Steam flooding

Not too difficult

Expensive
(x3 - x10)

Moderate
High in limited
geographical areas

Coats (1978), Prats (1982),


Mathews (1983)

Thermal Model In Situ


Combustion

In situ combustion
processes

Very difficult
Very specialised

Expensive
(x10 - x40)

Very low

Crookston et al (1979),
Youngren (1980),
Coats (1980b)

Institute of Petroleum Engineering, Heriot-Watt University

Amount of Industrial
Experience

Example References2

Simulator Type

as above

49

6 SOME FURTHER READING ON RESERVOIR SIMULATION


A full alphabetic list of References which are cited in the course is presented in
Appendix A. Here, we briefly review some good texts which cover Reservoir
Simulation from various viewpoints. The authors have learned something from each
of these and we would recommend anyone who wishes to specialise in Reservoir
Simulation to consult these.
Archer, J S and Wall, C: Petroleum Engineering: Principles and Practice, Graham
and Trotman Inc., London, 1986.
This book is not a specialised reservoir simulation text. However, it offers a good
overview of petroleum engineering and it contexts reservoir simulation very well
within the overall picture of reservoir development. This book is also one of the
earliest proponents of the importance of integrating the reservoir geology within the
simulation model.
Aziz, K. and Settari, A.: Petroleum Reservoir Simulation, Elsevier Applied Science
Publishers, Amsterdam, 1979.
This is a classic text on the discretisation and numerical solution of the reservoir
simulation flow equations. It is quite mathematical with a focus on the actual difference
equations that arise from the flow equations and how to solve these.
Crichlow, H B: Modern Reservoir Engineering: A Simulation Approach, PrenticeHall Inc., Englewood Cliffs, NJ, 1977.
This book gives a fairly good introduction to reservoir simulation from the viewpoint
of it being a central part of current reservoir engineering.
Dake, L P: The Practice of Reservoir Engineering, Developments in Petroleum
Science 36, Elsevier, 1994.
Again, this book is not about reservoir simulation but it makes a number of interesting
and controversial observations on reservoir simulation (not all of which the authors
agree with!). An interesting lengthy quote from this book on the relationship between
material balance and reservoir simulation is reproduced in Chapter 2.
Fanchi, J R: Principles of Applied Reservoir Simulation, Gulf Publishing Co.,
Houston, TX, 1997.
This recent book provides a good elementary text on reservoir simulation. It is a
based around the BOAST4D black oil simulation model which is supplied on disk and
can be run on your PC. The software makes this a very attractive way to familiarise
yourself with reservoir simulation if you dont have ready access to a simulator.
Mattax, C C and Dalton, R L: Reservoir Simulation, SPE Monograph, Vol. 13, 1990.
This is an excellent SPE monograph which covers virtually every aspect of traditional
reservoir simulation. It is has been put together by a team of Exxon reservoir engineers
between them have vast experience of all areas of reservoir simulation.
Peaceman, D W: Fundamentals of Numerical Reservoir Simulation, Developments
in Petroleum Science No. 6, Elsevier, 1977.
50

Introduction and Case Studies

This book presents an excellent treatment of the mathematical and numerical aspects
of reservoir simulation. It discusses the discretisation of the flow equations and the
subsequent numerical methods of solution in great detail.
SPE Reprint No. 11, Numerical Simulation I (1973) and SPE Reprint No. 20,
Numerical Simulation II (19**).
These two collections present some of the classic SPE papers on reservoir simulation.
All aspects of reservoir simulation are covered including numerical methods,
solution of linear equations, the modelling of wells and field applications. Most of
this material is too advanced or detailed for a newcomer to this field but the volumes
contain excellent reference material. They are also relatively cheap!
Thomas, G W: Principles of Hydrocarbon Reservoir Simulation, IHRDC, Boston,
1982. This short volume is written - according to Thomas - from a developers
viewpoint; i.e. someone who is involved with writing and supplying the simulators
themselves. The treatment is quite mathematical with quite a lot of coverage of
numerical methods. The treatment of some areas is rather brief; for example, there
are only 7 pages on wells.

APPENDIX A:

REFERENCES

NOTE: SPEJ = Society of Petroleum Engineers Journal - there was an early version
of this and it stopped for a while. Currently, there are SPE Journals in various
subjects but reservoir simulation R&D appears in SPE (Reservoir Engineering and
Evaluation).
Acs, G., Doleschall, S. and Farkas, E., General Purpose Compositional Model,
SPEJ, pp. 543 - 553, August 1985.
Allen, M.B., Behie, G.A. and Trangenstein, J.A.: Multiphase Flow in Porous Media:
Mechanics, Mathematics and Numerics, Lecture Notes in Engineering No. 34,
Springer-Verlag, 1988.
Amyx, J W, Bass, D M and Whiting, R L: Petroleum Reservoir Engineering, McGrawHill, 1960.
Appleyard, J.R. and Cheshire, I.M.: Nested Factorization, paper SPE 12264
presented at the Seventh SPE Symposium on Reservoir Simulation, San Francisco,
CA, November 16-18, 1983.
Archer, J S and Wall, C: Petroleum Engineering: Principles and Practice, Graham
and Trotman Inc., London, 1986.
Aronofsky, J.S. and Jenkins, R.: A Simplified Analysis of Unsteady Radial Gas
Flow, Trans., AIME 201 (1954) 149-154
Aziz, K. and Settari, A.: Petroleum Reservoir Simulation, Elsevier Applied Science
Publishers, Amsterdam, 1979.

Institute of Petroleum Engineering, Heriot-Watt University

51

Bondor, P.L., Hirasaki, G.J and Tham, M.J., Mathematical Simulation of Polymer
Flooding in Complex Reservoirs, SPEJ, pp. 369-382, October 1972.
Clayton, C.A., et al, The Ubit Field Rejuvenation: A Case History of Reservoir
Management of a Giant Oilfield Offshore Nigeria, SPE49165, presented at the SPE
Annual Technical Conference and Exhibition, New Orleans, LA, 27-30 September
1998.
Coats, K.H., .......... 1969 - tools of res sim
Coats, K.H., A Highly Implicit Steamflood Model, SPEJ, pp. 369-383, October
1978.
Coats, K.H., An Equation of State Compositional Model, SPEJ, pp. 363-376,
October 1980a; Trans. AIME, 269.
Coats, K.H., In-Situ Combustion Model, SPEJ, pp. 533-554, December 1980b;
Trans. AIME 269.
Coats, K.H., Dempsey, J.R., and Henderson, J.H.: The Use of Vertical Equilibrium
in Two-Dimensional Simulation of Three-Dimensional Reservoir Performance, Soc.
Pet. Eng. J. 11 (March 1971) 63-71; Trans., AIME 251
Craft, B C, Hawkins, M F and Terry, R E: Applied Petroleum Reservoir Engineering,
Prentice Hall, NJ, 1991.
Craig, F F: The Reservoir Engineering Aspects of Waterflooding, SPE monograph,
Dallas, TX, 1979.
Crichlow, H B: Modern Reservoir Engineering: A Simulation Approach, PrenticeHall Inc., Englewood Cliffs, NJ, 1977.
Crookston, R.B., Culham, W.E. and Chen, W.H., A Numerical Simulation Model for
Thermal Recovery Processes, SPEJ, pp. 35-57, February 1979; Trans. AIME 267.
Dake, L P: The Fundamentals of Reservoir Engineering, Developments in Petroleum
Science 8, Elsevier, 1978.
Dake, L P: The Practice of Reservoir Engineering, Developments in Petroleum
Science 36, Elsevier, 1994.
Fanchi, J R: Principles of Applied Reservoir Simulation, Gulf Publishing Co.,
Houston, TX, 1997.
Fantoft, S., Reservoir Management of the Oseberg Field After Four Years, SPE25008,
proceedings of the SPE European Petroleum Conference, Cannes, France, 16-18
November 1992.
Giudicelli, C.S., Massonat, G.J. and Alabert, F.G., Anguille Marine, a DeepseFan Reservoir Offshore Gabon: From Geology Toward History Matching Through
52

Introduction and Case Studies

Stochastic Modelling, SPE25006, proceedings of the SPE European Petroleum


Conference, Cannes, France, 16-18 November 1992.
Harpole, K.J. and Hearn, C.L., The Role of Numerical Simulation in Reservoir
Management of a West Texas Carbonate Reservoir, SPE10022, presented at the
International Petroleum Exhibition and Technical Symposium of the SPE, Beijing,
China, 18 - 26 March 1982.
Heinemann, Z.E., Brand, C.W., Munka, M., and Chen, Y.M.: Modeling Reservoir
Geometry with Irregular Grids, SPERE 6 (1991) 225-232.
Hove, K., Olsen, G., Nilsson, S., Tonnesen, M. and Hatloy, A., From Stochastic
Geological Description to Production Forecasting in Heterogeneous Layered Systems,
SPE24890, the proceedings of the SPE 67th Annual Technical Conference, Washington,
DC, 4-7 October 1992.
Katz, D.L., Methods of Estimating Oil and Gas Reserves, Trans. AIME, Vol. 118,
p.18, 1936 (classic early ref. on Material Balance)
Kyte, J.R. and Berry, D.W.: New Pseudo Functions to Control Numerical Dispersion,
Soc .Pet. Eng. J. 15 (August 1975) 269-276.
Lantz, R.B.: Quantitative Evaluation of Numerical Diffusion (Truncation Error),
Soc .Pet. Eng. J. 11 (September 1971) 315-320; Trans., AIME 251.
Leonard, A.J., Duncan, A.E., Johnson, D.A. and Murray, R.B., SPE25059:
Development Planning in a Complex Reservoir: Magnus Field UKCS Lower
Kimmeridge Clay Formation (LKCF), SPE25059, proceedings of the SPE European
Petroleum Conference, Cannes, France, 16-18 November 1992.
Mathews, C.W., Steamflooding, J. Pet. Tech., pp. 465-471, March 1983; Trans.
AIME 275.
MacDonald, R.C. and Coats, K.H.: Methods for Numerical Simulation of Water and
Gas Coning, Soc. Pet. Eng. J. 10 (December 1970) 425-436; Trans., AIME 249.
Mattax, C C and Dalton, R L: Reservoir Simulation, SPE Monograph, Vol. 13,
1990.
Meijerink, J.A. and Van der Vorst, H.A.: An Iterative Solution Method for Linear
Systems of Which the Coefficient Matrix is a Symmetric M-Matrix, Mathematics
of Computation 31 (January 1977) 148.
Nolen, J.S., Numerical Simulation of Compositional Phenomena in Petroleum
Reservoirs, SPE4274, proceedings of the SPE Symposium on Numerical Simulation
of Reservoir Performance, Houston, TX, 11-12 January 1973.
Palagi, C.L. and Aziz, K.: Use of Voronoi Grid in Reservoir Simulation, SPE
Advanced Technology Series 2 (April 1994) 69-77.
Peaceman, D.W.: Interpretation of Well-Block Pressures in Numerical Reservoir
Institute of Petroleum Engineering, Heriot-Watt University

53

Simulation, Soc. Pet. Eng. J. 18 (June 1978) 183-194; Trans., AIME 253.
Peaceman, D.W. and Rachford, H.H.: The Numerical Solution of Parabolic and
Elliptic Differential Equations, Soc Ind. Appl. Math. J. 3 (1955) 28-41
Peaceman, D W: Fundamentals of Numerical Reservoir Simulation, Developments
in Petroleum Science No. 6, Elsevier, 1977.
Ponting, D.K.: Corner point geometry in reservoir simulation, in The Mathematics
of Oil Recovery - Edited proceedings of an IMA/SPE Conference, Robinson College,
Cambridge, July 1989; Edited by P.R. King, Clarendon Press, Oxford, 1992.
Pope, G.A. and Nelson, R.C., A Chemical Flooding Compositional Simulator,
SPEF, pp.339-354, October 1978.
Prats, M., Thermal Recovery SPE Monograph Series No. 7, SPE Richardson, TX,
1982.
Price, H.S. and Coats, K.H.: Direct Methods in Reservoir Simulation, Soc. Pet.
Eng. J. 14 (June 1974) 295-308; Trans., AIME 257
Robertson, G., in Cores from the Northwest European Hydrocarbon Provence, C D
Oakman, J H Martin and P W M Corbett (eds.), Geological Society, London. 1997.
Schilthuis, R.J., Active Oil and Reservoir Energy, Trans. AIME, Vol. 118, p.3,
1936; (original ref. on Material Balance)
Sheldon, J.W., Harris, C.D., and Bavly, D.: A Method for Generalized Reservoir
Behavior Simulation on Digital Computers, SPE 1521-G presented at the 35th
Annual SPE Fall Meeting, Denver, Colorado, October 1960.
Sibley, M.J., Bent, J.V. and Davis, D.W., Reservoir Modelling and Simulation of
a Middle Eastern Carbonate Reservoir, SPE36540, proceedings of the SPE 71st
Annual Conference and Exhibition, Denver, CO, 6-9 October 1996.
Sorbie, K.S., Polymer Improved Oil Recovery, Blakie and SOns & CRC Press,
1991.
SPE Reprint No. 11, Numerical Simulation I (1973) and SPE Reprint No. 20,
Numerical Simulation II (19**).
Spillette, A.G., Hillestad, J.H., and Stone, H.L.: A High-Stability Sequential Solution
Approach to Reservoir Simulation, SPE 4542 presented at the 48th Annual Fall
Meeting of the Society of Petroleum Engineers of AIME, Les Vegas, Nevada,
September 30-October 3, 1973.
Stone, H.L.: Iterative Solution of Implicit Approximations of Multidimensional Partial
Differential Equations, SIAM J. Numer.Anal. 5 (September 1968) 530-558
Stone, H.L.: Probability Model for Estimating Three-Phase Relative Permeability,
J. Pet. Tech. 24 (February 1970) 214-218; Trans., AIME 249.
54

Introduction and Case Studies

Stone, H.L.: Estimation of Three-Phase Relative Permeability and Residual Oil


Data, J. Can. Pet. Tech. 12 (October-December 1973) 53-61.
Stone, H.L. and Garder, Jr., A.O.: Analysis of Gas-Cap or Dissolved-Gas Drive
Reservoirs, Soc .Pet. Eng. J. 1 (June 1961) 92-104; Trans., AIME 222.
Thomas G.W. and Thurnau, D.H.: Reservoir Simulation Using an Adaptive Implicit
Method, Soc. Pet. Eng.. J. 23 (October 1983) 759-768.
Thomas L.K., Lumpkin, W.B., and Reheis, G.M.: Reservoir Simulation of Variable
Bubble-point Problems, Soc. Pet. Eng. J. 16 (February 1976) 10-16; Trans., AIME
261.
Todd, M.R. and Longstaff, W.J.: The Development, Testing, and Application of a
Numerical Simulator for Predicting Miscible Flood Performance, J. Pet. Tech. 24
(July 1972) 874-882; Trans., AIME 253.
Todd, M.R., ODell, P.M., and Hirasaki, G.J.: Methods for Increased Accuracy in
Numerical Reservoir Simulators, Soc. Pet. Eng. J. 12 (December 1972) 515-530.
Thomas, G W: Principles of Hydrocarbon Reservoir Simulation, IHRDC, Boston,
1982.
Todd, M.R. and Chase, C.A., A Numerical Simulator for Predicting Chemical
Flood Performance, SPE7689, proceedings of the SPE Symposium on Reservoir
Simulation, Denver, CO, 1-2 February 1979.
Todd, M.R. et al , Numerical Simulation of Competing Chemical Flood Designs,
SPE7077, proceedings of the SPE Symposium on Improved Methods for Oil
Recovery, Tulsa, OK, 16-19 April 1978.
Uren, L.C., Petroleum Production Engineering, Oil Field Exploitation, 3rd edn.,
McGraw-Hill Book Company Inc., New York, 1953.
Van Quy, N. and Labrid, J., A Numerical Study of Chemical Flooding - Comparison
with Experiments, SPEJ, pp.461-474, June 1983; Trans. AIME 275.
van Vark, W., Paardekam, A.H.M., Brint, J.F., van Lieshout, J.B. and George, P.M.,
The Construction and Validation of a Numerical Model of a Reservoir Consisting
of Meandering Channels, SPE25057, proceedings of the SPE European Petroleum
Conference, Cannes, France, 16-18 November 1992.
Vela, S., Peaceman, D.W. and Sandvik, E.I., Evaluation of Polymer Flooding in a
Layered Reservoir with Crossflow, Retention and Degradation, SPEJ, pp. 82-96,
April 1976.
Vinsome, P.K.W.: Orthomin, an Iterative Method for Solving Sparse Banded Sets
of Simultaneous Linear Equations, paper SPE 5729 presented at the Fourth SPE
Symposium on Reservoir Simulation, Los Angeles, February 19-20, 1976.
Institute of Petroleum Engineering, Heriot-Watt University

55

Wallis, J.R., Kendall, R.P., and Little, T.E.: Constrained Residual Acceleration of
Conjugate Residual Methods, SPE 13536 presented at the Eighth SPE Reservoir
Simulation Symposium, Dallas, Texas, February 10-13, 1985.
Watts, J.W.: An Iterative Matrix Solution Method Suitable for Anisotropic Problems,
Soc Pet. Eng .J. 11 (March 1971) 47-51; Trans., AIME 251.
Watts, J.W., A Compositional Formulation of the Pressure and Saturation Equations,
SPE (Reservoir Engineering), pp. 243 - 252, March 1986.
Watts, J.W., Reservoir Simulation: Past, Present and Future, SPE Reservoir
Simulation Symposium, Dallas, TX, 5-7 June 1997.
Yanosik, J.L. and McCracken, T.A.: A Nine-Point Finite Difference Reservoir
Simulator for Realistic Prediction of Unfavorable Mobility Ratio Displacements,
Soc. Pet. Eng. J. 19 (August 1979) 253-262; Trans., AIME 267.
Young, L.C. and Stephenson, R.E., A Generalised Compositional Approach for
Reservoir Simulation, SPEJ, pp. 727-742, October 1983; Trans. AIME 275.
Youngren, G.K., Development and Application of an In-Situ Combustion Reservoir
Simulator, SPEJ, pp. 39-51, February 1980; Trans. AIME 269.

APPENDIX B - Some Overview Articles on Reservoir Simulation


1. Reservoir Simulation: is it worth the effort? SPE Review, London Section monthly
panel discussion November 1990.
This one pager summarises a panel discussion that was held in London in 1990. Given
the brevity of the article, it is packed with some genuine wisdom - and some things to
disagree with - from a really excellent group of front line users of the technology.
Briggs comes closest to capturing the principal authors particular prejudices!
2. The Future of Reservoir Simulation - C. Galas, J. Canadian Petroleum Technology,
36, January 1997.
This short viewpoint from a Canadian independent consultant is interesting since it
contexts reservoir simulation in the current outsourced and downsized oil industry.
He notes that virtually everyone can have PC based powerful simulation technology
on their desk tops. However, he concludes that the overall demand for simulation
will rise and that, for the sake of efficiency, this will be performed by specialists.
At the same time, he promotes a teamwork environment for the simulation engineer
where he or she will be involved in the preceding reservoir characterisation process
and the subsequent decision making process. This is a well argued position but not
all of his conclusion would be generally accepted.
3. What you should know about evaluating simulation results - M. Carlson; J.
Canadian Petroleum Technology, Part I - pp. 21-25, 36, No. 5, May 1997; Part II
- pp. 52-57, 36, No. 7, August 1997.
56

Introduction and Case Studies

This very interesting pair of articles gives a very good broad brush commentary
on a range of technical issues in reservoir simulation e.g. gridding, handling wells,
pseudo-relative permeability, error analysis and consistency checking. The views
are clearly those of someone who has been deeply involved in applied reservoir
simulation. They are well presented and quite individual although again there are
issues that would provoke disagreement. Read this and decide for yourself what you
accept and what you dont.

Institute of Petroleum Engineering, Heriot-Watt University

57

Basic Concepts in Reservoir Engineering

1.

INTRODUCTION

2.

MATERIAL BALANCE
2.1 Introduction to Material Balance (MB)
2.2 Derivation of Simplified Material
Balance Equations
2.3 Conditions for the Validity of Material
Balance

3.

SINGLE PHASE DARCY


LAW

3.1 The Basic Darcy Experiment


3.2 Mathematical Note: on the Operators
gradient and divergence
3.3 Darcys Law in 3D - Using Vector and
Tensor Notation
3.4 Simple Darcy Law with Gravity
3.5 The Radial Darcy Law

4.

TWO-PHASE FLOW
4.1 The Two-Phase Darcy Law

5.

CLOSING REMARKS

6.

SOME FURTHER READING ON


RESERVOIR ENGINEERING

LEARNING OBJECTIVES:
Having worked through this chapter the student should:
be familiar with the meaning and use of all the usual terms which appear in
reservoir engineering such as, Sw, So, Bo, Bw, Bg, Rso, Rsw, cw, co, cf, kro, krw, Pc etc.
be able to explain the differences between material balance and reservoir
simulation.
be aware of and be able to describe where it is more appropriate to use material
balance and where it is more appropriate to use reservoir simulation.
be able to use a simple given material balance equation for an undersaturated oil
reservoir (with no influx or production of water) in order to find the STOOIP.
know the conditions under which the material balance equations are valid.
be able to write down the single and two-phase Darcy Law in one dimension (1D)
and be able to explain all the terms which occur (no units conversion factors need
to be remembered).
be aware of the gradient () and divergence (.) operators as they apply to
the generalised (2D and 3D) Darcy Law (but these should not be committed to
memory).
know that pressure is a scalar and that the pressure distribution, P(x, y, z) is a
scalar field; but that P is a vector.
know that permeability is really a tensor quantity with some notion of what this
means physically (more in Chapter 7).
be able to write out the 2D and 3D Darcy Law with permeability as a full tensor
and know how this gives the more familiar Darcy Law in x, y and z directions when
the tensor is diagonal (but where we may have kx ky kz).
be able to write down and explain the radial Darcy Law and know that the pressure
profile near the well, P(r), varies logarithmically.

Basic Concepts in Reservoir Engineering

REVIEW OF BASIC CONCEPTS IN RESERVOIR ENGINEERING


Brief Description of Chapter 2

This module reviews some basic concepts of reservoir engineering that must be
familiar to the simulation engineer and which s/he should have covered already. We
start with Material Balance and the definition of the quantities which are necessary
to carry out such calculations: , co , cf , Bo , Swi etc. This is illustrated by a simple
calculator exercise which is to be carried out by the student. The same exercise is then
repeated on the reservoir simulator. Alternative approaches to material balance are
discussed briefly. The respective roles of Material Balance and Reservoir Simulation
are compared.
The unit then goes on to consider basic reservoir engineering associated with fluid
flow: the single phase Darcy law (k), tensor permeabilities, k , two phase Darcy Law
- relative permeabilitites (kro , krw) and capillary pressures (Pc).
Note that many of the terms and concepts reviewed in this section are summarised in
the Glossary at the front of this chapter.

1. INTRODUCTION

It is likely that you will have completed the introductory Reservoir Engineering part
of this Course. You should therefore be fairly familiar with the concepts reviewed
in this section. The purpose of doing any review of basic reservoir engineering is
as follows:
(i) Between them, the review in this section and the Glossary make this course more
self-contained, with all the main concepts we need close at hand;
(ii) This allows us to emphasise the complementary nature of conventional reservoir
engineering and reservoir simulation;
(iii) We would like to review some of the flow concepts (Darcys law etc.), in a
manner of particular use for the derivation of the flow equations later in this course
(in Chapter 5).
An example of point (ii) above concerns the complementary nature of Material Balance
(MB) and numerical reservoir simulation. At times, these have been presented as
almost opposing approaches to reservoir engineering. Nothing could be further from
the truth and this will be discussed in detail below. Indeed, a MB calculation will be
done by the student and the same calculation will be performed using the reservoir
simulator.
In addition to an introductory review of simple material balance calculations, we will
also go over some of the basic concepts of flow through porous media. These flow
concepts will be of direct use in deriving the reservoir simulation flow equations in
Chapter 5. Again, most of the concepts are summarised in the Glossary.
Exercises are provided at the end of this module which the student must carry out.

Institute of Petroleum Engineering, Heriot-Watt University

The following concepts are defined in the Glossary and should be familiar to you:
viscosity (o, w, g), density (o, w, g), phase saturations (So, Sw and Sg), initial or
connate water saturation (Swi or Swc), residual oil saturation (Sor). In addition, you
should also be familiar with the basic reservoir engineering quantities in Table 1
below:
Symbol

Name

Field Units Meaning / Formulae

Bo, Bw, Bg

Formation volume
factors (FVF) for oil,
water and gas

bbl/STB
or RB/STB

Bo = Vol. oil + dissolved gas in reservoir


Vol. oil at STC
STC = Stock Tank Conditions (60F; 14.7 psi).
Likewise for water (usually const.) and gas; Pb
= bubble point pressure below.

FVF

Bg

Bo
Bw
Pb

P
Rso, Rsw

Gas solubility factors


or solution gas oil
ratios

SCF/STB

Rso =

Vol. dissolved gas in reservoir


Vol. gas at STC

Rso
Rso
P
co, cw, cg

Isothermal fluid
compressibilities of
oil water and gas

Pb

-1

psi

ck = 1
k

k
P

=-

1
Vk

Vk
P

k and Vk - density and volume of phase k;


k = o, w, g

2. MATERIAL BALANCE
2.1 Introduction to Material Balance (MB)

The concept of Material Balance (MB) has a central position in the early history of
reservoir engineering. MB equations were originally derived by Schilthuis in 1936.
There are several excellent accounts of the MB equations and their application to
different reservoir situations in various textbooks (Amyx, Bass and Whiting, 1960;
Craft, Hawkins and Terry, 1991; Dake, 1978, 1994). For this reason, and because
this subject is covered in detail in the Reservoir Engineering course in this series,
we only present a very simple case of the material balance equation in a saturated
reservoir case. The full MB equation is presented in the Glossary for completeness.
Our objectives in this context are as follows:

To introduce the central idea of MB and apply it to a simple case which we


will then set up as an exercise for simulation;

Table 1: Basic reservoir


engineering quantities to
revise

Basic Concepts in Reservoir Engineering

To demonstrate the complementary nature of MB and reservoir simulation


calculations.
Material balance has been used in the industry for the following main purposes:
1. Determining the initial hydrocarbon in place (e.g. STOIIP) by analysing mean
reservoir pressure vs. production data;
2. Calculating water influx i.e. the degree to which a natural aquifer is supporting
the production (and hence slowing down the pressure decline);
3. Predicting mean reservoir pressure in the future, if a good match of the early
pressure decline is achieved and the correct reservoir recovery mechanism has been
identified.
Thus, MB is principally a tool which, if it can be applied successfully, defines the
input for a reservoir simulation model (from 1 and 2 above). Subsequently, the mean
field pressure decline as calculated in 3 above can be compared with the predictions
of the numerical reservoir simulation model.
Before deriving the restricted example of the MB equations, we quote the introduction
of Dakes (1994) chapter on material balance.

Material Balance Applied to Oilfields

(from Chapter 3; L. P. Dake, The Practice of Reservoir Engineering, Developments


in Petroleum Science 36, Elsevier, 1994.) Dake says:
It seems no longer fashionable to apply the concept of material balance to oilfields,
the belief being that it has now been superseded by the application of the more
modern technique of numerical reservoir simulation modelling. Acceptance of this
idea has been a tragedy and has robbed engineers of their most powerful tool for
investigating reservoirs and understanding their performance rather than imposing
their wills upon them, as is often the case when applying numerical simulation directly
in history matching.
As demonstrated in this chapter, by defining an average pressure decline trend for a
reservoir, which is always possible, irrespective of any lack of pressure equilibrium,
then material balance can be applied using simply the production and pressure
histories together with the fluid PVT properties. No geometrical considerations
(geological models) are involved, hence the material balance can be used to calculate
the hydrocarbons in place and define the drive mechanisms. In this respect, it is the
safest technique in the business since it is the minimum assumption route through
reservoir engineering. Conversely, the mere act of construction of a simulation
model, using the geological maps and petrophysically determined formation properties
implies that the STOIIP is known. Therefore, history matching by simulation can
hardly be regarded as an investigative technique but one that merely reflects the input
assumptions of the engineer performing the study.

Institute of Petroleum Engineering, Heriot-Watt University

There should be no competition between material balance and simulation, instead they
must be supportive of one another: the former defining the system which is then used
as input to the model. Material balance is excellent at history matching production
performance but has considerable disadvantages when it comes to prediction, which
is the domain of numerical simulation modelling.
Because engineers have drifted away from oilfield material balance in recent years,
the unfamiliarity breeds a lack of confidence in its meaningfulness and, indeed, how
to use it properly. To counter this, the chapter provides a comprehensive description of
various methods of application of the technique and included six fully worked exercises
illustrating the history matching of oilfields. It is perhaps worth commenting that in
none of these fields had the operators attempted to apply material balance, which
denied them vital information concerning the basic understanding of the physics of
reservoir performance.
Notes on Dakes comments
1. The authors of this Reservoir Simulation course would very much like to echo
Dakes sentiments. Performing large scale reservoir simulation studies does not
replace doing good conventional reservoir engineering analysis - especially MB
calculations. MB should always be carried out since, if you have enough data to
build a reservoir simulation model, you certainly have enough to perform a MB
calculation.
2. Note Dakes comments on the complementary nature of MB in defining the
input for reservoir simulation, as we discussed above.
3. Take careful note of Dakes comment on where a reservoir simulation model
is used for history matching. The very act of setting up the model means that you
actually input the STOIIP, whereas, this should be one of the history matching
parameters. The reservoir engineer can get around this to some extent by building
a number of alternative models of the reservoir and this is sometimes, but not
frequently, done.

2.2 Derivation of Simplified Material Balance Equations

Material balance (MB) is simply a volume balance on the changes that occur in the
reservoir. The volume of the original reservoir is assumed to be fixed. If this is so,
then the algebraic sum of all the volume changes in the reservoir of oil, free gas,
water and rock, must be zero. Physically, if oil is produced, then the remaining oil,
the other fluids and the rock must expand to fill the void space left by the produced
oil. As a consequence, the reservoir pressure will drop although this can be balances
if there is a water influx into the reservoir. The reservoir is assumed to be a tank
- as shown in Figure 5 Chapter 1. The pressure is taken to be constant throughout
this tank model and in all phases. Clearly, the system response depends on the
compressibilities of the various fluids (co, cw and cg) and on the reservoir rock formation
(crock). If there is a gas cap or production goes below the bubble point (Pb), then the
highly compressible gas dominates the system response. Typical ranges of fluid and
rock compressibilities are given in Table 2:

Basic Concepts in Reservoir Engineering

Table 2: Typical rock and


fluid compressibilities (from
Craft, Hawkins and Terry,
1991)

-6

Fluid or formation

Compressibility (10

Formation rock, crock


Water, cw
Undersaturated oil, co
Gas at 1000psi, cg
Gas at 5000psi. cg

3 - 10
2-4
5 - 100
900 - 1300
50 - 200

-1

psi )

The simple example which we will take in order to demonstrate the main idea of
material balance is shown in Figure 1 where the system is simply an undersaturated
oil, with possible water influx.
Initial conditions
pressure = po

Figure 1.
Simplified system for
material balance (MB) in
a system with an
undersaturated oil above the
bubble point and possible
water influx.

After production (Np)


pressure = p

Oil

Oil

(N - Np)Bo

NBoi = Vf.(1-Swi)

NBoi = Vf.(1-Swi)

Water, Swi

Water, Swi

W = Vf.Swi

W + We - Wp

Water influx

Oil, Np

Water, Wp

Water influx We

Definitions:
N
= initial reservoir volume (STB)
Boi = initial oil formation volume factor (bbl/STB or RB/STB)
Np = cumulative produced oil at time t, pressure p (STB)
Bo = oil formation volume factor at current t and p (bbl/STB)
W = initial reservoir water (bbl)
Wp = cumulative produced water (STB)
Bw = water formation volume factor (bbl/STB)
We = water influx into reservoir (bbl)
cw = water isothermal compressibility (psi-1)
P = change in reservoir pressure, p - po
Vf = initial void space (bbl); Vf = N.Boi/(1- Swi); W = Vf.Swi
Swi = initial water saturation (of whole system)
cf

= void space isothermal compressibility (psi-1); c f =

1 Vf

Vf p

(NB: (i) bbl = reservoir barrels, sometimes denoted RB; and (ii) in the figures
above, the oil and water are effectively assumed to be uniformly distributed
throughout the system)

Institute of Petroleum Engineering, Heriot-Watt University

Definitions of the various quantities we need for our simplified MB equation for the
depletion of an undersaturated oil reservoir above the bubble point (Pb) are given
in Figure 1. (NB a more extensive list of quantities required for a full material
balance equation in any type of oil or gas reservoir is given in the Glossary for
completeness).
In going from initial reservoir conditions shown in Figure 1 at pressure, po, to pressure,
p, volume changes in the oil, water and void space (rock) occur, Vo, Vw, Vvoid
(Vvoid = - Vrock). The pressure drop is denoted, P = p - po. The volume balance
simply says that:

Vo + Vw + Vrock = Vo + Vw Vvoid = 0

(1)

Each of these volume changes can be calculated quite


__ straightforwardly as
W
(W
W
B
+
W
+
.
.
p

V
=

W
c
follows:
w
p
w
e
w
Oil volume
= WVBo
Vchange,
w

__

- We - W.c w . p

Initial oil volume in reservoir

=
__

- (Vt,f pressure
- Vf .c f .p p )
= Vtf time
Oil volume

= (N - Np). Bo

__

= V .c . p

f volume,
f
Change in oil

N.Boi

(bbl = RB)
(bbl)

= N.Boi - (N - Np). Bo (bbl) (1)

Vo

__

Vrockchange,
Vvoid = - Vf .c f . p
= - V
Water volume
w
Initial reservoir water volume

= W

Cumulative water production at time = Wp


Reservoir volume of cumulative water production at time
= Wp.Bw
Volume of water influx into reservoir

= We

Water volume change due to compressibility


= W.cw. P
Change in water volume,

Vw

(bbl)
(STB)
(bbl)
(bbl)
(bbl)

= W - (W - Wp Bw + We + W.cw. P )
(bbl)

Vw

= Wp Bw - We - W.cw. P

(2)

= Vf

(bbl)

Change in the void space volume, Vvoid


Initial void space volume

Basic Concepts in Reservoir Engineering

Change in void space volume, Vvoid

= Vf - (Vf - Vf.cf. P )
= Vf.cf. P

Change in rock volume, Vrock = - Vvoid = - Vf.cf. P

(3)

Now adding the volume changes as follows:


__

__

Vo + Vw + Vrock = N . Boi + ( N N p ). Bo + Wp . Bw We W .cw . P Vf .c f . P = 0


(4)

Swi .cw + cf

__

__

__

+VN p . Bo=4N
Wand
P =V
0 = N.B /(1-S ), we
Rearranging
V .S .and
o
p .(B
.eB+oiW+noting
VNo. B+oiVNw . B+equation
Nw Nthat
N. Bpoi).W
Bo 1+=W
rock
Swipf . Bwiw We f W .cw .oiP Vwif .c f . P = 0
obtain:

Swi .cw + c f __ S .c + cf __
N. BW
. P = 0 wi w
p . Bo. B
oi + W . B
NN..BBoioi NN. B
. Bo o++NN
. P = 0
p
o
e 1 pS w N . Boi
wi
1 Swi
(5)
Swi .cw + crock __
__
__
N . Boi N . Bo + N p . Bo + N . Boi S .c + .c P = 0__
wi
wbalance
fN ).
.
(

=
+

+
.

.
.

.
.

P
V
V
V
N
B
N
B
W
B
W
W
c
P
V
c
1

Equation
5
is
the
(simplified)
material
expression
for
the
undersaturated
wi
o
w
oi
p
o
p
w
e
w
rock
f
f
. P = 0
N.B N.B + N .B N.B
oi

oi

Swi above its bubble point).


system given in Figure 1 (as longas it1
remains
S .c + c

__

w
rock
Swiexample,
.cw + cf let __
To Nillustrate
in an
us assume
. Peven
= 0 simpler
. Boi N . Bthe
N p . Bof
N . Boi wibalance
o + use
o +material
N
.
B

N
.
B
+
N
.
B

W
+
W
.
B
N

.
B

S
1

. P =the0 MB
oi water oinfluxp (W
o =0)
e wi
pc w
oi = 0). Therefore,
__
that there is no
or
production
(W
S
c
.
+
wi w
rock
e
p 1 Swi
+ N . Boito:
N . Boi simplifies
N . Bo + Neven

. P = 0
p . Bo further
equation
1 S

wi

S .c + c __
N . Boi N . Bo + N p . Bo N . Boi wi w __f . P = 0
S .c + c S
. P=0 __
N . Boi N . Bo + N p . Bo + N . Boi wi w S 1rock
wi
(6)
wi .cw+ crock
1

N
B
N
B
N
B
N
B
.

.
+
.
+
.
.

P
=
0
wi

oi
o
p
o
oi
Note that we can divide through equation
(the
6Sby.cN +
__reservoir oil volume,
c initial
N .obtain:
Boi N . Bo + N p . Bo + N . Boi wi1 wSwi rock . P = 0
bbl = RB) to
S .c1 + Scwirock __
N . Boi N . Bo + N p . Bo + N . Boi wi w
. P = 0
1 Swi

__
Np
Swi .cw + c f
Boi Bo + N . Bo Boi S .c + c . P__= 0
(7)
Boi Bo +N p . Bo Boi 1wi Swwi f . P = 0
N
1 Swi .c + crock __
N . Boi N
. Bo +to:N p . Bo + N . Boi wi w
. P = 0
which rearranges
easily
1 __Swi
Boi Boi Swi .cw + c f
Np
+
N = 1 B
Swi .cw + cf . P__= 0
N p = 1 Bo oi +BB
1 Swi . P = 0
oi
o

(8)
N
Bo Bo 1 Swi
where the __quantity (Np/N) is the Recovery Factor (RF) as a fraction of the STOIIP.
p__= 0
It is seen from
equation 8 that, at t = 0, Bo = Boi and and therefore (Np/N)= 0, as
p=0
expected. Note also in equation 8 that P is negative in depletion ( P = p-po,
where po. is the
k for
P depletion).
Q higherkinitial
P pressure

u= = .
= .
AQ k LP k xP
u=
= .
= .
A
L
x

Institute of Petroleum Engineering, Heriot-Watt University

It is convenient to rearrange equation 8 above as follows:

N p Boi Boi Swi cw + c f

1
=

N Bo Bo 1 Swi

(9)

We then identify 1-(Np/N) as the fraction of the initial oil still in place. We can then
plot this quantity vs. - P shown in Figure 2 (we take - P since it plots along the
positive axis, since P is negative).
"almost" straight line
for w/o systems

1
1-

Np
N

Figure 2
Plot of remaining oil,

Np

1
vs. P

-P

As noted in Figure 2, this decline plot is not necessarily a straight line but for oil water
systems, it is very close in practice. Figure 2 suggests a way of applying a simple
material balance equation to the case of an undersaturated oil above the bubble point
(with no water influx or production). This is a pure depletion problem driven by the
oil (mainly), water and formation compressibilities. Suppose we know the pressure
behaviour of B0 (i.e. B0(P)) as shown in Figure 3.
1.4

Bo(P) = m.P + c

Oil FVF
Bo
1.3

4000

P (psi)

5500

Figure 3
B0 as a function of pressure
for a black oil.

If we draw the reservoir pressure down by an amount P (known or measured) and


we know that to do this we had to produce a volume Np (STB) of oil. This point of
depletion is shown in Figure 4.
1
1-

Np
N
Y
0

10

-P

Figure 4
Reservoir depletion on a
plot following equation 9.

Basic Concepts in Reservoir Engineering

We know Y ( it is P ), we can calculate X (the RHS of equation 9). X is equal to 1(Np/N) and we know Np (the amount of oil we had to produce to get drawdown P ).
Hence, we can find N the initial oil in place. An exercise to do this is given below.

2.3 Conditions for the Validity of Material Balance

The basic premise for the material balance assumptions to be correct is that the
reservoir be tank like i.e. the whole system is at the same pressure and, as the
pressure falls, then the system equilibrates immediately. For this to be correct, the
pressure communication through the system must at least be very fast in practice
(rather than instantaneous which is strictly impossible). For a pressure disturbance
to travel very quickly through a system, we know that the permeability should
be very high and the fluid compressibility should be low (pressure changes a re
communicated instantaneously through and incompressible fluid). Indeed, we will
show later (Chapter 5) that pressure equilibrates faster - or diffuses through the
system faster - for larger values of the hydraulic diffusivity, which is given by
k/(c) (Dake, 1994, p.78).
Dake (1994, p.78), also points out two necessary conditions to apply material
balance in practice as follows:
(i) We must have adequate data collection (production/pressures/PVT); and
(ii) we must have the ability to define an average pressure decline trend i.e. the more
tank like, the better and this is equivalent to having a large k/(c) as discussed
above.

EXERCISE 1.
Material Balance problem for an undersaturated reservoir using equation 8 above.
This describes a case where production is by oil, water and formation expansion
above the bubble point (Pb) with no water influx or production.
Exercise:
Suppose you have a tank - like reservoir with the fluid properties given below (and
in Figure 4). Plot a figure of

Np

1
vs. - P over the first 250 psi of depletion

of this reservoir. Suppose you find that after 200 psi of depletion, you have
produced 320 MSTB of oil. What was the original oil in place in this reservoir?
Input data: The initial water saturation, Swi = 0.1. The rock and water
compressibilities are, as follows:
cf = 5 x 10-6 psi-1;

cw = 4 x 10-6 psi-1.

The initial reservoir pressure is 5500 psi at which Boi = 1.3 and the bubble point is
at Pb = 4000 psi where Bo = 1.4. That is, the oil swells as the pressure drops as
shown in Figure 4.

Institute of Petroleum Engineering, Heriot-Watt University

11

3. SINGLE PHASE DARCY LAW


We review the single phase Darcy Law in this section in order to put our own particular
slant or viewpoint to the student. This will prove to be very useful when we derive
the flow equations of reservoir simulation in Chapter 5. We also wish to extend
the idea of permeability (k) somewhat further than is covered in basic reservoir
engineering
we wish to introduce the idea of permeability as
texts. In particular,
a tensor property, denoted by k . Some useful mathematical concepts will also be
introduced in this section associated with vector calculus; in particular, the idea of
gradient and divergence will be discussed in the context of the generalised
formulation of the single phase Darcy law. Note
for reference,
__ many of the terms
Swi .cthat
w + crock
N . Boiare also
N . Bsummarised
N .the
Boi Glossary.
. P = 0
o + N p . Bo + in
discussed here
1 S

wi

3.1 The Basic Darcy Experiment


Darcy in 1856 conducted
through
__ packs of sands which he took
S .ctests
N a series of flow
w + cf
.

0 ground water supply at


Boi wiof an
Boi experimental
Bo + p . Bo models
as approximate
aquifer
for=the
P
1 Sexperiment

wi
Dijon. A schematic ofNthe essential Darcy
is shown in Figure 5 where

we imagine a single phase fluid (e.g. water) being pumped through a homogeneous
sand pack or rock core. (Darcy used a gravitational head of water as his driving force
__
c f normally
B
B S .cw +would
p
whereas, inNmodern
= 1 coreoilaboratories,
+ oi wi we
. P = 0 use a pump.)

Bo

Bo 1 Swi

The Darcy law given in Figure 5, is in its experimental form where a conversion
factor, , is indicated that allows us to work in various units as may be convenient
__
to the problem
p =at0hand. In differential form, a more useful way to express the Darcy
Law and introducing the Darcy velocity, u, is as follows:

k P
k P
Q
u= = .
= .
A
L
x

(9)

where the minus sign in equation 9 indicates that the direction of fluid flow is down
the pressure gradient from high pressure to low pressure i.e. in the opposite direction
to the positive pressure gradient.
P

Q
L
Q = .

k.A P
.

Definitions:
Symbol Dimensions

Meaning

Consistent Units
c.g.s
lab.

field

SI field

L3/T

Volumetric
flow rate

cm3/s

cm3/s

bbl/day

m3/day

Length of
system

cm

cm

ft.

L2

Cross - sectional cm2

cm2

ft.2

m2

12
A

Q
Basic Concepts in Reservoir Engineering
L
Q = .

k.A P
.

Definitions:
Symbol Dimensions

Consistent Units
c.g.s
lab.

field

SI field

L3/T

Volumetric
flow rate

cm3/s

cm3/s

bbl/day

m3/day

Length of
system

cm

cm

ft.

L2

Cross - sectional cm2


area

cm2

ft.2

m2

Viscosity

cP

cP

cP

Pa.s

Figure 5.
The single phase Darcy Law

Meaning

M.L.T.2
(Force/Area)

Pressure drop

atm

dyne/cm2

psi

Pa

L2

Permeability#

darcy

darcy

mD

mD

dimensionless Conversion
factor

1.00

9.869x10

-6

-3

1.127x10

8.527
x10-3

# permeability - dimensions L2; e.g. units m2, Darcies (D), milliDarcies (mD); 1 Darcy
= 9.869 x 10-9 cm2 = 0.98696 x 10-12 m2 1 m2.

Note on Units Conversion for Darcys Law: the various units that are commonly used
for Darcys Law are listed in Figure 2 above. Sometimes, the conversion between
various systems of units causes confusion for some students. Here, we briefly explain
how to do this using
in the previous figure; that is, we go from c.g.s.
k.A the
examples
P
Q -=gram
. - second)
. units where = 1, indeed, the Darcy was defined such
(centimetre
L
that = 1. Starting
k.Afrom
Pthe
Darcy Law in c.g.s. units:

Q = .

.
L

k (Darcy) . A (cm 2 ) P (atm )


.

(cp)
2 L ( cm )
k (Darcy) . A (cm ) P (atm )
.
Q (cm 3 / s) = 1.00

L (cm )
(cp)
Suppose we now wish to convert to field units as follows:
k (Darcy) . A (ft 2 ) P ( psi)
Q ( bbl / day) = ??
.

(cp)
2 L ( ft.)
k (Darcy) . A (ft ) P ( psi)
Q ( bbl / day) = ??
.

L (ft.)
(cp)
k
P
( mD) . A (ft.2 ).30.482
( psi)
bbl 1.58999x10 5

1000 for these new units? Essentially,


.7we
How do we
Q find the
. correct conversion
. 14
= factor
4
k

P
2
2

8starting
.64 x10 from
cp
30
L (ft=.).(1.
day
( mD). (A
()ft. ).we
.48 that
30know
psi.48
)
5 the c.g.s. expression
convert it unit
by unit
where
bbl
1.58999
x10

1000
14
.
7

Q to know
. a few conversion
. (exact),
We do need
= factors as follows: 1 ft. = 30.48 cm
4

8.64
day 1 bbl
3(cp
)
30.48
x10
14.7 psi = 1 atm.,
= 5.615
ft3= 5.615 x 30.483 cm
= 1.58999
x 105 cm
day
L3,(1ft.).

= 24 x 3600 s = 8.64 x 104 s. Thus, we now


4 convert
2 everything in the field units to2

.64which
bbl
x10 . 30.48
c.g.s. units
for 8cp.
are the same): k ( Darcy) . A ( ft. ) . P( p
Qas follows

= (except
L (ft
(cp)
day 1000 . 1.58999x410 5. 14.27 . 30.48
bbl
k (Darcy) . A (ft.2 ) P( p
8.64 x10 . 30.48
Q
.

=
L (ft
(cp)
day 1000 . 1.58999x10 5. 14.7 . 30.48
2
bblEngineering,

k (Darcy) . A (ft. ) P( psi)


Institute of Petroleum
Heriot-Watt
13
-3 University
Q
.

= 1.126722x10
(cp)
day
2 L ( ft.)
P psi

Q (cm 3 / s) = 1.00

k.Ak.AP P
Q=
Q=. 3. . L. k (Darcy) . A (cm 2 ) P (atm )
L

Q (cm / s) = 1.00

(cp)

L (cm )

k (Darcy) . A (cm 2 ) 2P (atm )


. ) P (atm )
k (Darcy) . A 2(cm
(cp
) )
Q (cm / s) = 1.00
( psi

k (Darcy
) ). A (ft ) LP(.cm

Q (cm 3 / s3) = 1.00

(cp)
(cp)

Q ( bbl / day) = ??

. L (cm
)
L (ft.)

k (Darcy) . A (ft 2 ) P ( psi)


2
k (Darcy) . A .(ft L
) (ft
P ( psi)
Q ( bbl / day) = ?? (cp)k
. .) 2

Q ( bbl / day) = ??

P
(Lft.(ft)..)30.482
(cp)( mD) . A
( psi)
bbl 1.58999x10

1000
14
.
7
Q
.
.
= k
4

((ft
cp.2)).30.482 PL (psi
ft.).30
day 8.64 x10

.48
(
)
.
mD
A
(
)
5
bbl 1.58999x10

1000 k
142.7 P
2
Q
.
. .48
=5
4
(
)
.
(
.
).
mD
A
ft
30
psi
(
)

64 x10 x10
(cp)
30.48
day

bbl 8.1.58999

L (ft.).
Q
.
= 1000
. 14.7

Thus, collecting
thex10
numerical
day 8.64
(cpobtain:
)
L (ft.).30.48
factors togetherwe
bbl
k (Darcy) . A (ft.2 ) P( psi)
8.64 x10 4. 30.482
Q
.

=
5
L (ft.)
(cp)2
day 1000 . 1.58999
4 x10 . 14
2 .7 . 30.48
5

bbl
k (Darcy) . A (ft. ) P( psi)
8.64 x10 . 30.48
Q
.

=
5

L (2ft.)
day
1000
.
1.58999
x
10
14
7
30
48
.
.
.
.
(cp)

4
2

bbl
k (Darcy) . A (ft. ) P( psi)
8.64 x10 . 30.48
2
Q
.
P( psi)

=
5
bbl

(
)
.
(
.
)
k
Darcy
A
ft
-3

(
.)
L
ft
1000
.
1.58999
x10 . 14.7 . 30.48

cp
(
)

Q day
=
1.126722
x
10
.

which
simplifies
to
2

day

cp
(
)
(
.)
L
ft
bbl

-3 k ( Darcy ) . A ( ft. ) P( psi )


Q
= 1.126722x10
day

(cp)

L2 (ft.)

bbl
-3 k ( Darcy ) . A ( ft. ) P( psi )
Q
.

= 1.126722x10
L (ft.)
(cp)
day

and hence = 1.127 x 10-3 for these units (as given in Figure 5).

3.2 Mathematical Note: on the Operators gradient and


divergence

Before generalising the Darcy Law to 3D, we first make a short mathematical digression
to introduce the concepts of gradient and divergence operators. These will be used
to write the generalised flow equation of single and two phase flow in Chapter 5.
Gradient (or grad) is a vector operation as follows:

=
= x ii +
+ y jj +
+ z kk
= x i + y j + z k
x
y
z
where i, ij, and
k are the unit vectors which point in the x, y and z directions,
i , jj and
and kk
respectively.
The
operation can be carried out on a scalar field such as
i , j and gradient
k
pressure, P, as follows:

P
P i + P
P j + P
P
P

=
P
i
+
j + Pz kk

y
P = x i + y j + z k
x
y
z

where P
is sometimes written as grad P. The quantity P is actually a vector of

P
P
the pressure
Pgradients in the three directions, x, y and z as follows:

P
P i
P
x i
xx i
P j
P

P
P =
= P
j
P = yy j
y
P k
P
Pzz kk
z
14

P
P
P
P
P ii,, P jj,, and
and kk

P
i
x

Basic
P Concepts

P =
j
y

P k
z

in Reservoir Engineering

This is shown schematically in Figure 6 where the three components of the vector

P
P, i.e. i,
x

P
P
j, and k , and are shown by the dashed lines.
z
y

Figure 3: The definition of grad P or

Unit vectors
k

j
y

Figure 6
The definition of grad P or
P

Divergence (or div) is the dot product of the gradient operator and acts on a vector
to produce a scalar. The operator is denoted as follows:

.
.
.
.
.
.
. .
.
.. .
. ..
.. .
. ..
. .

. .
. .
. .
. . ..
. .
. .
.
.. .
. .
.. ..
..
.
.

.
.
.


= i
j
k
y
z
x
= i
j
k
y
z
x
For example, taking
velocity vector, u, gives the
the divergence

of the Darcy

u
i
x

=
i
j
k
following:


z i j
u =x i y j =
k
k uu yx ij

j
= i
k
z
zx y
y x z y
x

u = i
j
k u zy kj

y
z u i
x
x

u z k
ux i
u = i
j u x i k u y j

u
i
y u =
z i x
x

ju
k uu y j

=
u
i
j
k
u
j

u
y the

u
k

y equation
where wecan
x z
by
y multiplying
u = the
k x uu zxy ij =
i i + out
j j +first z k k
y i RHSzofj the
xexpand
above

matrix
(1x3) matrix by the second
which
x (3x1)
y matrix
zkobtain
a 1x1
uzto
uyuyisz kascalar

ux
uzz
u = i
j
k u zy kj = x i i +
jj +
k k
as follows:
y
z u i x
y
z
x

x
uuy i

u z k u x

u
i i =u j =j = k ik = 1 j u x i k u y j =
i i + xj j + uz k k
y u =
z u xu y k yuujz = zx i i +
x

u = i
j
k u y j =xui z kxi iy+j
j j + y k k x

i ix= j j =ky k = 1 z
z
y
x
u z k
u z k

u y
u x
u z
i i =u j =j = k k =+1
+
ux
uy
uz

y
where
i i += j j z= k k = 1 , to obtain:
uk =relationships,
i i =we
j juse
= kthe
= 1 x +

y
z

u y
u z
P Pu x
u =
+
+
u y
u z
z u x
u x Puxy
uy
u =P
+
+ z u = x + y + z

y
z
x
P
i
P P
x
Institute of Petroleum Engineering, Heriot-Watt University
15
P

P P i
P P
2
2

P

P
x P

.
.

.
.

.. . .
.
.
. . .
.
.
.
.
.
.
.
.
.
. . . .. .

. .

.
.
.
.

. u = x i.

. u j = ux i.i +

j
y

. . .

u y

i i = j j = k k = 1 u z k

. . .

i i = j j = k k =1

k
z

.. .

. u = ux

u y
u z
+
y = zx i

u y
u
u z
u = x +
+
y
z
x

jj +

u z
k k
z


k
z

j
y

uquantity,

Likewise, we can take the divergence of the grad P vector, P, to obtain the
xi

P P

P, (sometimes denoted div grad


P),
= follows:
u as
i
j
k uy j

. .

. . . .

. .
.

. . .. .

. .

.
.

i i = j j = k k = 1 , we obtain the familiar

. P =. xPP += yP P + + zP.P=u+=P uP += u P +

x y

. x
z
y

. . .

=j1jrelationships,
i i = j using
j =i ki k=the
Again
= k k =1
expression:

z
x P y

i
u z k
x
P
i

P 2 P

2 P
2 P

P = i jP k 2 P j =2 P 2 i i2 uP+ i 2 j j + 2 k k
xk
k y j z= iyi + 2
j xj +
P = i
zu
jx
ky
u z

y z 2 u x
y
z y x 2
x
y

u = i
j
k
u
j
=
+
+
i
i
j
j
k k

P
y
P x
z
y
z

x
k
k
u
k

z
z

y
2

u z
z

where, in summary, 2 is the Laplacian


P operator:
P

2 =

. = div . grad = x

2
2
+
y 2
y 2

P
i
x
The final issue we wish to discuss in this mathematical note is the rule for
taking
now
P of 2 P
2 P
2 P
omitthe explicit
inclusion
the dot product of a tensor and a vector. N.B. We
k xx k xy k xz

P
=
i
j
k
j
=
+
+
i
i
j
j

2 k k

the unit vectors,


z
y
z y x 2
x
y 2
i, j and k in the
following developments.

k = k yx k yy k yz
P
k in

2 and Tensor
2
2
3.3 Darcys
3D k- zzUsing
Vector
Notation
2 Law
k
k
zx
zy
2
2
2
=have
atensor
= div k grad
= 23D+ can
+ 2
a 3zx 3
2be represented

Suppose we
which in
by
2
= = div grad = x 2 + y 2 + y 2
matrix as follows:
x
y
y
P k P
i i = j j = k k =1
k xx k xy k xz
k
k
k xz
k = k xx
k yz
k xy
yx
yy
P 2
k = k yx k yy k yz
2 P
2 P 2
P

k
k
k

P
=
+
+

x
zx
zy
zz
2 = P
2
2

x
z
y
k xx k zzk xy k xz

k zx kzy

P
k product
k
= k
k Pwish

of
Suppose we
Pnow
k P toyxtake yya dot yz
y this tensor, k , with the vector P;

that is k
product of a tensor and a vector is a vector and the operation
Pdot
P P. kThe
k zx k zy k zz P
is carried out like a matrix multiplicationas follows:


Pz

P
x

k xx k xy k xz x
k
k xz P P
k
P
P
k yz P
k xy
k P = k xx
k xx + k xy + k xz
yx
yy
P

z
y
k P = k yx k yy k yz yx x
k
k
k

zx
zy
zz
kk xx kk xy kk xz P

zx
zy
zz
P
P
P
P

P
k P = k yx k yy k yz z = k yx + k yy + k yz
y x
z
y
k

zx k zy k zz P

k zx P + k zy P + k zz P
P
P
Pz
P
16
k xx xP + k xy yP + k xz Pz

P
x k xx x + k xy y + k xz z

..

. .
.
..
.

.
.

. . .
.

kPP k P

Basic Concepts
Engineering
P in
P
Reservoir

x x
k xx kxyk k xz k

xx

xy P xz
P

k
k yx
k
k Pk =

P = yyk yx yz kyy yk yz
k

zx kzyk k zz k Pk
zx
zy
zz

P
z
which multiplies out as follows:
z

P k P + k P + k P

xx
xy
xz
x P x k Py + k Pz + k P


xx
xy
xz
k xx k xy k xz

x
z
y

P x P

k yx + k yy P + k yz P
P = k yx kyyk xxk yz k
xy k xz =

P x Py
Pz
P
k zx
k
P =kzyk yxk zz kyy k yz = k yx + k yy + k yz
P y

x
z
y
k
k k zx P + k zy P + k zz P
k

zy z zz
zx
z
y
P x
P
P
P

k zx
x

+ k zy
y

giving the final


result:
P
P
P
k xx + k xy + k xz
z
y
x

P
Pk xx P+ k xy P + k xz

z
k P = k yx + kxyy + kyzy

x
z
y

P P P
P P P
+ k
k
P = k
+ k
k zx yx+ kxzy yy+ kzzy yz z
x
z
y

+ k zz
z

P
k zx
x

P
+ k zy
y

P
+ k zz
z

Using the above concepts from vector calculus (div. and grad), we can extend the
Darcy Law (in the absence of gravity) to 3D as follows by introducing the tensor
permeability, k :

u= -

1
k
u

P
P
PP
P
k xz +
k xx P+ k xy +k xx

x
xz
xx y
k xx k xy k xz

P
kxx
1
P k xy1 k xzP
P
P = 1- k yx k yy k yz1 = - k yx P+ k yy 1 + k yz P
= - kk kP = k- kyxy k yy k yzx = -y k yx z +

x
zx zy zz P
Py P
P

k zx k zy kkzz
+ k zy + k zz
zx
z

which we maywrite
Pas:

P
P
k xx + k xy + k xz
ux

P P P P

1 P
k xx + k+yz k xy
k yy
u = u y = - kuyx +
z y
x x

y x

uz
k zy P + k zz P
k zx 1+

x u = u y =
k y
+ kzyy
yx



1 P u
u x = - k xx z+

x
y

P
P
PP
k xy k+zx k xz + k zy

z
y
y
x

P
1 P
P
u y = - k yx + k yy + k yz
z P
x 1
yP
u =- k
+ k

xx Heriot-Watt
+
x
xy
Institute of Petroleum
Engineering,
University
1

k zx
x

P
k yy
y

P
+ k zy
y

P
+ k xz
z

P
+ k yz
z

P
+ k zz
z

P
k xz
z

P
k xy
y

17

x
P
P
P
k
+
k
+
k

xx
xy
xz
ux
x
z
y

P
P
P

k
+
k
+
k

xx
xy
xz
u x
Pz
x
y
P
1 P
u = u y = - k yx + k yy + k yz
z
x
y

1
P
P
P
u = u y = - k yx + k yy + k yz

zP

Px
Py
uz
k
+
k
+
k


zx

zy
zz


P
velocity
P
P
and we can identify
components
of
the
as
follows:
u z the three
k zx + k zy + k zz
z
y
x

P
1 P
P
u x = - k xx + k xy + k xz
z
x
y
1
P
P
P
u x = - k xx + k xy + k xz
z
x
y
P
1 P
P
u y = - k yx + k yy + k yz
z
x
y
1
P
P
P
u y = - k yx + k yy + k yz
z
x
y
P
1 P
P
u z = - k zx + k zy + k zz
z
x
y
1
P
P
P
u z = - k zx + k zy + k zz
z
x
y
If the permeability tensor is diagonal i.e. the cross-terms are zero as follows:
k xx 0
0

k = 0k
k0yy 00
xx
0

k = 0
k0yy k0zz
0
0
k zz

then the various components of the Darcy law revert to their normal form and :

1
P
k xx
x

1
P
u x = - 1 k xx P
1 u = -P k xx x
u x = - kxxx
x
u y = -1x k yy P

y
P
1
u y = - 1 k yy P
1 u = -P k yy y
u y = - kyyy
1 Py
u z = -y k zz
z

1
P
u z = - 1 k zz P
1 u = -
P k zz z with Gravity
3.4
- kzzz Darcy
u z =Simple
zP
1 Law
z
u x = -of
z gravity
gDarcy
k xx the-1D
Law becomes:
In the presence
x

x
1
z
P
u x = - 1 k xx P - g z
1 u = -P k xz - g x
u x = - kxxxz -xxgx
x
=x cos x
x
zcase
of a simple inclines system at a slope of , as shown in Figure 7,
where, inthe
xz = cos

z , =ascos
x 1 shown
= cos
Pin the figure above
and:
x u x = - k xx
- g.. cos
x

1
P

u x = - 1 k xx P - g.. cos

1 u = -P k x - g.. cos
u x = - kxxx 2 khr-xxgdP
..xcos


Q = x
dr
2 khr dP
18
Q = 2 khr dP
Q = dP dr
2 khr
ux = -

Basic Concepts in Reservoir Engineering

Note that:
z
x

= cos

Figure 7
Radial form of the singlephase Darcy Law

3.5 The Radial Darcy Law

In the above discussion, in both 1D and 3D we considered the Darcy Law in normal
Cartesian coordinates (x, y and z). In Chapter 6, we will explain how wells are
treated in reservoir simulation. Because a radial (r/z) geometry is appropriate for
the near-well region, it is useful to consider the Darcy Law in radial coordinates. In
1D, this simply involves the radial coordinate, r. In fact, the radial form of the Darcy
law can be derived from the linear form as shown in Figure 8.

1
P
k xx
P

1
xQ
u x = - k xx
x

ux = -

P
1
k yy

y
P
1
=
u
k

yy
h y

dr

uy = -

Figure 8
Single phase Darcy Law in
an inclines system - effect of
gravity

Area, A = 2.r.h

Radial Darcy Law is:

1
P
u z = - k zz
2khr dP
k.A dP
Q=
=
Pz

dr
dr
1

u z = - k zz
z

1Q
z flow rate of fluid into well
Notation:
=Pvolumetric
- g
u x = - k xx
r
radial distance
=P

xz from well
1
x
- g of formation
u x = - h k xx = height

x pressure drop from r (r + dr) i.e. over dr


dP =xincremental
= area of surface at r = 2.r.h
z A
= cos = fluid viscosity
xz
k
= cos = formation permeability
x r
= wellbore radius
w

1dr
=Pincremental radius

u x = - k xx
- g.. cos

x
1radial form
P

Starting from
the
of
the
Darcy
u x = - k xx
- g.. cos Law, as follows:

x
2 khr dP
Q=
dr
2 khr dP
Q=
dr
we can rearrangethis
Q todrobtain:


2Q
kh dr
r
dP =
2 kh r
dP =

Institute of Petroleum Engineering, Heriot-Watt University

19

Taking rw as the wellbore radius and r some appropriate radial distance, we can easily
integrate the above equation to obtain the radial pressure profile in a radial system
as follows:
r

Q
dr
= Q r dr
r dP

r
rw
rw dP = 22kh
kh rw r
w
r

which gives:

Q r
P( r ) = Q ln r
P( r ) = 2 kh ln rw
2 kh rw

Pw
z
k.k rw the
where weuhave
denoted
- gpressure
w z drop (or increase for a producer) from rw to
Pradial
w = - k.k
xw the
rw unlike
x Law, the pressure profile is logarithmic
r as, P(r).
u wNote
= - that,
- linear
gw Darcy
means
x that pressure
x drops are much higher closer to the well.
in the radial case. This
This is exactly what we expect physically since the area is decreasing with r as we
k.k ro QPiso the same;ztherefore,

approach uthe=well
the pressure drop, dP, over a given
- k.kand
P - go z
o
ro
o

dr is higher.
This
is
shown
schematically
for
an
injector
and a producer in Figure 9.
u o = - x - go x

x developed
x here will be used later in Chapter 4 on well
The formulae and the ideas
modelling in reservoir simulation and we will not discuss this further here.
Pw
P
P and Po
xw and xQo
x
x Injector

Q
Producer

Pwf

Pc (Sw ) = Po Pw
Pc (Sw ) =P(r)
Po =PP
wfw- P(r)

P(r)

P(r)
P(r) = P(r) - Pwf

Pc (Sw ) = Pnon wett . Pwett .


Pc (Sw ) = Pnon wett . Pwett .
k rww = k k rwr
k w = k k rw
4. TWO-PHASE
k o = k kFLOW
ro
k o = k k ro
4.1 The Two-Phase Darcy Law

Pwf
rw

Darcys Law was originally applied to single phase flow only. However, in reservoir
k w and k o
engineering,
it hask been convenient to extend it to describe the flows of multiple
k w and
phases such as oil, owater and gas. To do this, the Darcy Law has been modified
empirically to include a term - the relative permeability - which is intended to describe
the impairment of the flow of one phase due to the presence of another. A schematic
representation of a steady-state two phase Darcy type (relative permeability) experiment
is shown in Figure 10, where all of the quantities are defined. Examples of the relative
permeability curves which can be measured in this way are also shown schematically
in Figure 10 and actual experimental examples are given for rock curves of different
wettability states in the Glossary.

20

Figure 9
Pressure profiles, P(r), in
radial single-phase flow; Pwf
is the well flowing pressure
(at rw)

Basic Concepts in Reservoir Engineering

At steady - state flow conditions, the oil and water flow rates in and out,
Qo and Qw, are the same:

Po

Pw

Qw
Qo

Qw
Qo

The two - phase Darcy Law is as follows:

Qw =

Schematic of relative
permeabilities, krw and kro

k.krw.A Pw
.
L
w

Figure 10
The two-phase Darcy Law
and relative permeability

kro

Rel.
Perm.

k.kro.A Po
Qo =
.
L
o

krw

Sw

Where:
Qw and Qo
A
L
w and o
k
Pw and Po

= volumetric flow rates of water and oil;


= cross-sectional area;
= system length;
= water and oil viscosities;
= absolute permeabilities;
= the pressure drops across the water and oil phases at
steady-state
flow conditions
r
r
the
Q water
drand

r
r
krw and kdP
=
oil relative permeabilities
=
Q dr
ro

2 kh rw r

rw

NB the Units for the two-phase Darcy Law


5.

dP = 2kh r

arerwexactly

rw
the same
as those in Figure

P( r ) =form of ln
Q including
r
phase Darcy Law
The differential
gravity
ln
P( r )in= 1D, again
2 kh the rtwo
w
which is taken to act in the z-direction, is as follows: 2 kh rw
uw = -

z
k.k rw Pw
- gw

x
x

uw = -

z
k.k rw Pw
- gw

x
x

uo = -

z
k.k ro Po
- go

x
x

uo = -

z
k.k ro Po
- go

x
x

where we note
P that the flow
P of the two phases (water and oil, in this case) depends

and o
P
P
on the pressure
x gradient
inx that
phase; i.e. on w and o .
x
x
Pc (Sw ) = Po Pw

Pc (Sw ) = Po Pw

Institute of Petroleum Engineering, Heriot-Watt University

P (S

)=P

21

= Q dr
dP
k P
dP =2k.kh
r

rw

khrwrw wr - gw z
urw w = - 2 rw
x
x

Q r r
P(Pr()r=) k=.kQln

z
ln
ro P
u o = - 2 2kh
orwr-w go

kh
x
x

uw = -

z
k.k rw Pw
- gw

x
x

uo = -

z
k.k ro Po
- go

x
x

zz
k.k PwP
u wu =P=-w - k.rwk rw P
w- gw

o
g

Pand
w
wx
x

SPw (S
= 1 - SPo), are generally
and

The phase pressures,


P
,
at
a
given
x saturation,
w oand
w
x o xx w
x pressure,
x as
follows:
not equal. However, they are related through the capillary
zz
k.kk.rok PoP

u oP
co(S
u=
=- w-)= Proo x Po w- -ggoo x
x
x

Pc (Sw ) = Po Pw

More strictly, the capillary pressure is the difference between the non-wetting
P S = P P wett . Pwett .
Pcw(Pw)andnon
phase pressure
the
pressure; Pc (Sw ) = Pnon wett . Pwett .. We
Po
owetting-phase
of the
w and
and x pressure

can think
as a constraint on the phase pressures. That is, if
x capillary

x
x
we know kthe capillary
pressure function - from experiment , say - then, if we have
w = k k rw
k rw pressure curves
k w =of kcapillary
Po at a given saturation, we can calculate Pw. Examples
are alsoPshown
Glossary.
(SwS) =in=Pthe
cP
oP
PwP
c

( w)

k o = k k ro

Note that, as in the single-phase Darcy Law, we maykgeneralise


the two-phase Darcy
o = k k ro
expressions
3D.
Defining
the
combination
of
absolute
permeability
in its full
=

PcP(Swto
P
P
)
S = non
P wett . wett
P.

( )

wett .
wett .
c
kw, with non
tensor form,
k w and
k o the phase relative permeabilities gives:

k wk = =k kk rwk
w
rw
k ok = =k kk rok
o

k w and k o

ro

k o are the effective phase permeability tensors of water and oil,


where k wk and
w and k o
respectively. Using this notation, the Darcy velocity vectors for the water and oil, uw
and uo, may be written in 3D as follows:
uw =

1
k w .(Pw w gz)
w

uo =

1
k o .(Po o gz)
o

This form of these equations is particularly useful in deriving the two-phase flow
equations in their most general form (this will done in Chapter 5).

5. CLOSING REMARKS
The purpose of Chapter 2 is to review some key concepts in reservoir engineering
which impact directly on the subject matter of reservoir simulation. The topics
reviewed specifically involved:
- Material balance and its particular relationship with reservoir simulation;
- The single-phase Darcy law and its extension using vector calculus terminology
to a 3D version of the Darcy Law including tensor permeabilities;

22

Basic Concepts in Reservoir Engineering

- The two-phase Darcy Law and the related concepts that arise in two-phase
flow e.g. relative permeabilities (kro and krw), phase pressures (Po and Pw),
capillary pressure (Pc(Sw) = Po - Pw), etc.
Ideas and concepts developed here will be used in other parts of this course.

6. SOME FURTHER READING ON RESERVOIR ENGINEERING


A full alphabetic list of References which are cited in the course is presented in
Appendix A. Many excellent texts have appeared over the years covering the basics
of Reservoir Engineering. Some of these are listed below, although this list is far
from comprehensive.
Amyx, J W, Bass, D M and Whiting, R L: Petroleum Reservoir Engineering, McGrawHill, 1960. This is still an excellent petroleum engineering text although the coverage in
some areas a little old fashioned. It has a very good chapter on material balance.
Archer, J S and Wall, C: Petroleum Engineering: Principles and Practice, Graham
and Trotman Inc., London, 1986. This book offers a good overview of petroleum
engineering and covers many of the basics of reservoir engineering. This book is
also one of the earliest proponents of the importance of integrating the geology within
the reservoir model.
Craft, B C, Hawkins, M F and Terry, R E: Applied Petroleum Reservoir Engineering,
Prentice Hall, NJ, 1991. The original text by Craft and Hawkins was already an
early classic. This was revised and updated by Terry and reissued in 1991. This has
very good clear coverage of material balance and its application in various reservoir
systems.
Craig, F F: The Reservoir Engineering Aspects of Waterflooding, SPE monograph,
Dallas, TX, 1979. This text is confined to the underlying principles and reservoir
engineering applications of waterflooding. It is an excellent monograph on the
subject and an essential reference text for the reservoir engineer who is interested in
the traditional analytical methods for assessing waterflooding.
Dake, L P: The Fundamentals of Reservoir Engineering, Developments in Petroleum
Science 8, Elsevier, 1978. This has become a modern classic on the basics of reservoir
engineering. It is very widely referenced and draws on Dakes vast experience of
teaching reservoir engineering basics. It has particularly good coverage of material
balance and Buckley-Leverett theory.
Dake, L P: The Practice of Reservoir Engineering, Developments in Petroleum
Science 36, Elsevier, 1994. This book is a modern plea for the continued application
traditional reservoir engineering principles and techniques in performance analysis and
prediction. It gives central place to the interpretation of well testing, the application
of material balance and the use of Buckley Leverett theory. It has many examples
from the hundreds of reservoirs that Dake himself worked on. This book also makes
a number of interesting and controversial observations on reservoir simulation (not
all of which the authors agree with!).
Institute of Petroleum Engineering, Heriot-Watt University

23

Solution To Exercises
EXERCISE 1:
Material Balance problem for an undersaturated reservoir using equation 8 above.
This describes a case where production is by oil, water and formation expansion
above the bubble point (Pb) with no water influx or production.
Exercise: For the input data below, do the following:
(i) Plot the function (1 - N/Np) as calculated by equation 8 vs. -P.

As a reminder equation 8 is 1

Np
B
B Swi + c f
= 1 oi + oi
P
N
Bo Bo 1 Swi

This is shown below


(1-Np/N) vs. -DP
0.999
0.997

(1-Np/N)

0.995
0.993

Series 1

0.991
0.989
0.987
0.985
0

50

100

150

200

250

300

-p (psi)

(ii) Note from the graph (or from your numerical calculation when plotting the
graph) that, at - P = 200 psi, then (1 - Np/N) = 0.991. However, we know by field
observation that this 200 psi drawdown was caused by the production of 320 MSTB.
That is, we know that Np = 320 MSTB. Hence,
(1 - 320/N) = 0.991 => N = 35555.5 MSTB 35.6 MMSTB
Answer: the STOOIP = 35.6 MMSTB.
Input data: The initial water saturation, Swi = 0.1. The rock and water compressibilities
are, as follows:
crock = 5 x 10-6 psi-1;

cw = 4 x 10-6 psi-1.

The initial reservoir pressure is 5500 psi at which Boi = 1.3 and the bubble point is
at Pb = 4000 where Bo = 1.4. That is, the oil swells as the pressure drops as shown
below:

24

Basic Concepts in Reservoir Engineering

1.4

Bo(P) = m.P + c

Oil FVF
Bo
1.3

4000

P (psi)

Institute of Petroleum Engineering, Heriot-Watt University

5500

25

26

Reservoir Simulation Model Set-Up

CONTENTS
1.

INTRODUCTION TO CHAPTER 3

2. SETTING UP A RESERVOIR SIMULATION


MODEL
2.1. Defining The Objectives Of A Simulation
Study

3. DATA INPUT AND OUTPUT


4.

EXAMPLE INPUT DATA FILE


4.1. Reservoir System to be Modelled
4.2. ECLIPSE Syntax
4.3. Model Dimensions
4.4. Grid and Rock Properties
4.5. Fluid Properties
4.6. Initial Conditions
4.7. Output Requirements
4.8. Production Schedule

5.

RUNNING ECLIPSE AND FILE NAME


CONVENTIONS
5.1. Running ECLIPSE on a PC
5.2. File Name Conventions

6.

CLOSING REMARKS

LEARNING OBJECTIVES
Having worked through this chapter and the associated tutorials the student
should be able to:
Simulation Input
Identify what questions the simulation is expected to address.
Identify what data is required as input to perform the desired calculations.
Format data correctly, taking account of keyword syntax and required units.
Simulation Output
Select required output of calculations.
Quality check output data to check for errors in input.
Identify purpose of each output file and use post-processors to analyse data.
Analysis of Results
Identify impact of reservoir engineering principles in calculation performed.
Identify numerical effects and impact of grid block size and orientation on
results.
Perform simple upscaling calculation to address numerical diffusion.

Reservoir Simulation Model Set-Up

BRIEF DESCRIPTION OF CHAPTER 3


In this module, a step-by-step approach is given to setting up a 3D reservoir
simulation model. This is done by working through an actual case which is complex
enough to demonstrate most of the basic ideas, can demonstrate various sensitivities
and can also show the effects of well controls. The simulator used in this course is
ECLIPSE which is a software product of Schlumberger GeoQuest. However, the
general approach and methodology for setting up a field simulation calculation is
very similar for other commercial simulators.
This example will be used to illustrate the power of reservoir simulation in
understanding reservoir recovery mechanisms.

1. INTRODUCTION TO CHAPTER 3
In this section of the course, we set up a practical 3D, two phase (oil/water) reservoir
simulation model using the ECLIPSE reservoir simulator. This is proprietary
software of Schlumberger GeoQuest. The central objective of this exercise is to
get you actually applying reservoir simulation to a realistic (but quite simple) case.
However, there are also some tasks in the study itself - you can think of these as
the objectives if this were a real field case. One of the tasks in the exercise is
as follows: in your calculation, you should observe initial short-term rise in BHP
(bottom hole pressure) in the injection well and drop in BHP in the production well.
You are asked to explain these trends. This is good example of where you may
have run a calculation without necessarily thinking about what was going to happen
to the BHP (or it could be watercut at the producer, or field average pressure etc.).
However, when you study the simulator output - usually as graphs and figures - you
would notice the BHP trends. This would catch the attention of a good engineer
who would not be happy just to note it and move on. She or he would immediately
stop and think and ask a few questions, Whats going on here?, Is this something
physical that I should expect or is there something wrong with the calculation?.
The engineer would stop and work it out from their basic reservoir engineering
knowledge ... just as you are going to! The engineer would conclude that although I possibly didnt expect it - this behaviour is perfectly understandable
and predictable.
From the above discussion, you can see that it is not just the mechanics of running
a numerical simulator and getting the results out that we want you to achieve in
this course. We want you to be able to formulate the right questions for a given
reservoir application, carry out the appropriate simulations and then interpret the
results correctly. The mechanics of running a simulation - if this was all you did - is
really a technicians job, the important job of correctly formulating the simulation
problem, understanding the results and predicting reservoir performance is an
engineers job and this course is intended for the latter (or for the former strongly
intent on becoming the latter in the future!).

Institute of Petroleum Engineering, Heriot-Watt University

2. SETTING UP A RESERVOIR SIMULATION MODEL


This section will not be very long since there are many excellent examples of
reservoir simulation studies elsewhere in this course e.g. Cases 1 - 3 in Chaper 1, the
SPE field cases. Some generalities on how to set up a reservoir simulation study were
also discussed in Chaper 1. Here, we will lay out more formally, the general procedues
broadly following the workflow of a typical simulation study.

2.1 Defining The Objectives Of A Simulation Study

Defining the objectives is a vitally important stage of any field simulation study. The
general spirit which is suggested for approaching this (in Chaper 1) is to correctly
formulate the question you are trying to ask in order to make a particular decision.
For example, the decision may be: do I need to infill drill in this field in order to
significantly (i.e. economically) improve reservoir performance?. This is like the
schematic example in Chaper 1, Figure 8 where the question was not will I get
more oil by ..., since you could get more oil but at too great a cost - the decision
must be economically based. Having said this, some reservoir decisions are made that
may not in themselves be economic; however, they may be strategic or may lead to
some knowledge or experience which will be economic in the future. The important
matter in that you know what sort of decision you are trying to make.

3. DATA INPUT AND OUTPUT


When run, every reservoir simulator will require input data that defines the system
to be modelled, and should generate output data that represents the results of the
calculations which have been performed. Although different reservoir simulation
codes have different formats for entering data, they all have some basic components
in common. Most will read data from an input file. The input file must therefore
be set up before starting the simulation run. The data file may be set up by
manual editing (if it is in ASCII format), or by using a Graphical User Interface
(GUI). Whichever method is used, most data files will be divided into certain
key sections that define:
Model dimensions
Grid and rock properties
Fluid properties
Initial conditions
Output requirements
Production schedule
Additional optional sections may allow for manipulation of an imported grid
structure and for subdivision of the grid into regions. We will find that there are a
huge number of possible refinements in all of the above general sections representing
special models for particular applications but we will focus on the simpler common
features of most black oil simulations.
Individual parts of the input data may be set up by other programs that may be supplied
by the same supplier as the simulation code, or by other companies. These are referred
to as pre-processors: They are used to perform calculations that set up the model in

Reservoir Simulation Model Set-Up

advance of the actual reservoir fluid flow calculation. Typically, a pre-processor is


used to set up the grid and to define the rock properties (permeability, porosity, etc.)
of each cell. Setting up a model with more than a few hundred cells would be very
laborious if performed by hand. These gridding packages are usually designed to
generate output that conforms to the input format of several of the more widely used
simulators. Thus, the reservoir simulation input data file need only refer to these
grid files by means of a simple include statement, and no further manipulation of
the grid is required by the flow simulator.
Pre-processors may be used to:

Define grid and rock properties

Define fluid properties

Convert the results of special core analysis data to a form that can be
used in the simulation

Upscale rock data so that it is appropriate for the size of grid cells being used

Define vertical flow performance tables

Set up the production schedule


Indeed, any software that is used for setting up a part or the whole of an input
data file is termed a pre-processor.
The output of the reservoir flow calculations usually comes in two forms, which in
both cases results in the creation of files that can be stored and read at a later date. The
first category of output data is typically referred to as summary data and the second
type of output consists of grid data. The two types are as follows:
(a) Summary data: this consists of calculated parameters such as oil, water
and gas production rates, well bottom hole or tubing head pressures, etc. These
data may be plotted as line charts, usually as a function of time, either by using
specialised post-processing software, or by using standard graphing software such
as Microsoft Excel or Lotus 1,2,3.
(b) Grid data: in this type of data, values such as pressure or saturation to be
plotted for each cell at a given time step. These files are typically in binary format,
which means that they may only be read by appropriate post-processing software.
The reason that ASCII format is not generally used is one of disk space usage.
For example, a 100,000 cell model, with output data for 20 time steps, would
generate 2 million values of pressure (usually to eight significant figures)
during the course of the simulation, and similarly for every other property
such as phase saturations, etc.
Two major, and usually understated, elements of good reservoir simulation
practice are thus:
Keeping a record of what each calculation represents (by choosing sensible
file names and inserting comments)
Minimising disk usage (by outputting only data that is actually required).
Current software developments are addressing automatic report generation and
minimising the time taken to obtain a good history match by automatically
varying specified parameters.
Institute of Petroleum Engineering, Heriot-Watt University

4. EXAMPLE INPUT DATA FILE


4.1 Reservoir System to be Modelled

In the first tutorial exercise, Tutorial 1A, the reservoir system to be modelled
consists of a five-spot pattern of four production wells surrounding each injection
well. However, the symmetry of the system allows us to model a single
injection production pair, which will be located at opposite corners of a grid,
as shown in Figure 1.
The system is initially at connate water saturation (Swc), and a waterflood
calculation is to be performed to evaluate oil production and water breakthrough
time. The reservoir will be maintained above the bubble point pressure (Pb)
at all times, and thus there is no need to perform calculations for a free gas
phase (in the reservoir).
Reservoir and fluid properties, such as layer permeabilities, porosity, oil and water
PVT and relative permeability data, are provided, as are the initial reservoir conditions
and production schedule (proposed injection and production rates).
The input data file, whether generated using a text editor or by GUI, consists of
various sections that incorporate all of these components just described. Here we will
go through the input file, TUT1A.DATA, used for this calculation. TUT1A.DATA
will also be used as a base case for other tutorial sessions associated with this
course. The data format is that required by the Schlumberger GeoQuest Reservoir
Technologies model, ECLIPSE 100, but other than syntactical differences, the style
of data entry is similar for most other simulators.
While the data file may be set up using a GUI, it is useful in the first instance to
set up a simple model using a text editor, thus ensuring by the end of the exercise
that every line of data is familiar and relatively well understood. This file can then
be used as a starting point for other models, which may be set up by modifying
the appropriate parts of this data file.

Production well
Injection well
Two well quarter
five-spot grid

Figure 1.
A five spot pattern consists
of alternating rows of
production and injection
wells. The symmetry of the
system means that the flow
between any two wells can
be modelled by placing the
wells at opposite corners of
a Cartesian grid, and is
referred to as a quarter fivespot calculation.

Reservoir Simulation Model Set-Up

4.2 ECLIPSE Syntax

Each item of data, such as porosities or relative permeabilities, are identified by use
of set keywords. The individual sections are also designated by keywords. The
syntax, format and function of each keyword may be found in the online manuals.
These also give examples of how the keywords may be used.
There are certain rules governing data entry for any simulator, and effort
must be made at the outset to get these right; otherwise setting up the input
data file can be very frustrating and as time consuming as performing the
calculations themselves.
ECLIPSE uses free format. This means that, with a few exceptions, as many
or as few spaces, tabs and new lines may be used as desired. However,
arranging the file appropriately, such as by lining data up in columns, etc., can
improve readability, reducing unnecessary typographical mistakes, and saving
time in the long run.
The following additional rules should be noted
Each section starts with a keyword
There must be no other characters (or spaces) on the same line as a keyword
(i.e. each keyword must start in column 1, and be immediately followed
by a new line keystroke)
All data associated with a keyword must appear on the subsequent lines
Data entry is terminated by a forward slash symbol (/)
Lines beginning with two dashes (--) are ignored, and treated as comment
lines
Blank lines are ignored
To illustrate the use of keywords, data and comments, the following style conventions,
illustrated in Figure 2, will be used here.
Figure 2.
Example of conventions
used to identify components
of input data file. It should
be noted that the actual
input file should be in
ASCII text format only
(as produced by Notepad,
WordPad or other basic
text editor), and should
not contain italic, bold or
coloured letters.

FEATURE

EXAMPLE FROM DATA FILE

Comments
NX

KEYWORDS
Data (followed by /)

Number of cells
NY
NZ

DIMENS
5

3/

4.3 Model Dimensions

The first step in setting up a model is to define:


Title of run
Type of geometry to be used (Cartesian or radial, though Cartesian is
often the default)
Number of cells in each direction (x, y, z, or r, , z)
Phases to be modelled (oil, water, gas, vapourised oil in the gas, dissolved
gas in the oil)
Institute of Petroleum Engineering, Heriot-Watt University

Units to be used (field, metric or lab)


Number of wells
Start date for simulation (usually corresponding to the date of first oil
production)

The model set up in Tutorial 1A consists of a Cartesian grid of 5 x 5 x 3 cells, each cell
having dimensions of 500 ft x 500 ft x 50 ft, as shown in Figure 3.
500

500

500

500

500

1
2
3

50
50
50
1

500

500

500

500
500

Cartesian is the default geometry used by ECLIPSE, so it does need to be specified


explicitly. A two-phase oil and water calculation is to be performed in this model, with
the injection and production wells to be located at opposite corners, and completed in
all three layers. ECLIPSE allows wells to be grouped together so that the cumulative
production or injection rates may be specified or calculated. Here, we will assume
that the injection and production wells are in two separate groups. Field units are
to be used throughout the input data file. (Note that once a choice of units has been
made, it must be used consistently for all data entry. This precludes, for example,
using feet (field units) for depths and bars (metric units) for pressures in the same
run.) The title for this calculation will be 3D 2-Phase, and it is assumed that first
oil was on 1st January 2001. Generated output should be written to a single unified
output file. (The ECLIPSE default is to create a separate output file for every time
step, which has the advantage that not all the data output data is lost if one file is in
some way corrupted, but this may result in an unmanageable number of files being
generated.)
The above information is all that is required for defining the dimensioning data
that goes in the first section of an ECLIPSE data file, referred to as the RUNSPEC
section. The form in which this data should be entered is shown in Figure 4.
The following keywords are used:
RUNSPEC
DIMENS
OIL
WATER
FIELD
WELLDIMS
8

Section header
Number of cells in X, Y and Z directions
Calculate oil flows
Calculate water flows
Use field units throughout (i.e. feet, psi, lb, bbl, etc.)
Number of wells, connections per well, groups, wells per group

Figure 3.
Cartesian grid of 5 x 5 x
3 cells used to represent
reservoir system to be
modelled in Tutorial 1A.

Reservoir Simulation Model Set-Up

UNIFOUT
START

Unified output file


Start date of simulation (1st day of production)

Every time an unfamiliar keyword is encountered, it is well worth looking it up in


the online manual, and this is probably as good a point to start as any! Particular
attention should be paid to units. For example, when using field units gas rates
are entered in MSCF/day. Entering a value in SCF/day would be allowed by the
simulator, but would lead to completely wrong results.
TUT1A. DATA
Base case for tutorials

RUNSPEC
TITLE
3D 2-Phase
Number of cells
NX
NY
NZ
DIMENS
5

3/

Phases
OIL
WATER

Units
FIELD
Maximum well / connection / group values
#wells
#cons/w
#grps
#wells/grp
WELLDIMS
2

1/

Unified output files


UNIFOUT

Figure 4.
RUNSPEC Section of input
data file.

Simulation start date


START
1 JAN 2001 /

Institute of Petroleum Engineering, Heriot-Watt University

4.4 Grid and Rock Properties

Having identified the number of grid cells in the X, Y and Z directions required
to model the reservoir or part of the reservoir being studied, the following grid
properties must be defined:
Dimensions of each cell
Depth of each cell (or at least the top layer)
Cell permeabilities in each direction (x, y, z, or r, , z)
Cell porosities
If a Cartesian grid is being used, as here, then the size of each cell may be specified by
providing data on the length, width and height of each cell. The current grid has 75
cells (5 x 5 x 3), and thus 75 values must be specified for each property.
Each cell in the model is to be 500 ft long, by 500 ft wide, by 50 ft thick. There are
three layers, the top layer being at a depth of 8,000 ft. All three layers are assumed
to be continuous in the vertical direction, so there is no need to specify the
depths of the second and third layers - the simulator can calculate these implicitly
from the depth of the top layer and the thickness of the top and middle layers.
The formation has a uniform porosity of 0.25, and the layer permeabilities in
each direction are given below.

Layer
1
2
3

Permeability (mD)
Horizontal
Vertical
X direction Y direction Z direction
200
150
20
1000
800
100
200
150
20

This represents the minimum information that is required for defining the grid and
rock properties for the second section of an ECLIPSE data file, referred to as the GRID
Section. It is useful to output a file that allows these values to be viewed graphically
by one of the post-processors. This enables a quick visual check that the grid data has
been entered correctly. The following keywords are used:
GRID
DX
DY
DZ
TOPS
PERMX
PERMY
PERMZ
PORO
INIT

Section header
Size of cells in the X direction
Size of cells in the Y direction
Size of cells in the Z direction
Depth of cells
Cell permeabilities in the X direction
Cell permeabilities in the Y direction
Cell permeabilities in the Z direction
Cell porosities
Output grid values to .INIT file

ECLIPSE normally assumes that grid values, such as DY, DZ, PERMX, PORO, etc.,
are being entered for the whole grid. If values are only being entered for a subsection
of the grid, then the BOX and ENDBOX keywords may be used to identify this
subsection (an example is given later). If no BOX is defined, or after an ENDBOX
10

Reservoir Simulation Model Set-Up

keyword, ECLIPSE assumes that cell values are being defined for the entire grid.
The values of each grid property, such as cell length, DX, are read in a certain order.
If the co-ordinates of each cell are specified by indices (i, j, k), where i is in the X
direction, j is in the Y direction, and k is in the Z direction, then the values are read
in with i varying fastest, and k slowest. The first value that is read in is for cell
(1, 1, 1), and the last one is for cell (NX, NY, NZ), where NX, NY and NZ are the
number of cells in the X, Y and Z directions respectively.
Thus, in this (NX=5, NY=5, NZ=3) model, the values of DX (and every other
grid value such as DY, DZ, PERMX, PORO, etc.) will be read in in the order
shown in Figure 5.

Figure 5.
Order in which cell property
values are read in by
ECLIPSE, starting at (1, 1,
1), and finishing at (5, 5, 3),
with the i index varying the
fastest, and the k index the
slowest.

No
1
2
3
4
5
6
7
8
.
.
24
25
26
27
.
.
75

i
1
2
3
4
5
1
2
3
.
.
4
5
1
2
.
.
5

j
1
1
1
1
1
2
2
2
.
.
5
5
1
1
.
.
5

500

k
1
1
1
1
1
1
1
1
.
.
1
1
2
2
.
.
3

500

500

500

500

1
2
3

50
50
50
1

500

500

500

500

500

X
Z

The length (DX) of each of the 75 cells in Tutorial 1A is the same: 500 ft. Thus
the data may be entered as:
DX
500
500
500
500
500

500
500
500
500
500

500
500
500
500
500

500
500
500
500
500

500
500
500
500
500

500
500
500
500
500

500
500
500
500
500

500
500
500
500
/

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

500
500
500
500

Most simulators will allow the definition of multiple cells, each with the same
size, to be lumped together. In ECLIPSE this is done by prefixing the value
(cell size) by the number of cells to be assigned that value, and separating these
two numbers by a *. Thus, since all 75 cells in the model have a length of
500 ft, this may be entered as:
DX
75*500 /
Note that the multiplier comes first, then the * operator, then the value. There
should be no spaces on either side of the *.
Institute of Petroleum Engineering, Heriot-Watt University

11

This convention may be used for all other grid parameters.


If cell depths are only to be defined for the top layer of cells using the TOPS
keyword, then a box must be used to identify this top layer as the only section of the
grid for which depths are being defined. The box that encompasses the top layer is
defined as from 1 to 5 in the X direction, 1 to 5 in the Y direction, but only 1 in the Z
direction. Instead of 75 cells for the whole model, there are only 25 cells in this section
of the model, and thus only 25 values of TOPS need be defined:
BOX

1 /

TOPS
25*8000 /
ENDBOX
The GRID Section of the Tutorial 1A input data file should be as shown in
Figure 6.

12

Reservoir Simulation Model Set-Up

GRID
Size of each cell in X, Y and Z directions
DX
75*500 /

DY
75*500 /

DZ
75*50
TVDSS of top layer only
X1 X2
Y1

Y2

Z1

Z2

1/

BOX
1

TOPS
25*8000 /

ENDBOX
Permeability in X, Y and Z directions for each cell
PERMX
25*200 25*1000 25*200 /
PERMY
25*150 25*800 25*150 /
PERMZ
25*20 25*100 25*20 /
Porosity of each cell
PORO
75*0.2 /
Figure 6.
GRID Section of input data
file.

Output file with geometry and rock properties (.INIT)


INIT

Institute of Petroleum Engineering, Heriot-Watt University

13

4.5 Fluid Properties

Having defined the grid and rock properties such as permeability and porosity, the
following Pressure/Volume/Temperature (PVT), viscosity, relative permeability and
capillary pressure data must be defined:
Densities of oil, water and gas at surface conditions
Formation factor and viscosity of oil vs. pressure
Pressure, formation factor, compressibility and viscosity of water
Rock compressibility
Water and oil relative permeabilities, and oil-water capillary pressure
vs. water saturation
The densities of the three phases oil, water and gas are given below. (Note that all three
densities must be supplied, even though free gas is not modelled in the system.)
Oil
(lb/ft3)
49

Water
(lb/ft3)
63

Gas
(lb/ft3)
0.01

The oil formation volume factor (Bo) and viscosity (o) is provided as a function
of pressure (P).
Pressure
(psia)
300
800
6000

Oil FVF
(rb/stb)
1.25
1.20
1.15

Oil Viscosity
(cP)
1.0
1.1
2.0

At a pressure of 4,500 psia, the water formation volume factor (Bw) is 1.02 rb/stb,
the compressibility (cw) is 3 x 10-6 PSI-1 and the viscosity (w) is 0.8 cP. Water
compressibility does not change with pressure within the pressure ranges encountered
in the reservoir, and thus viscosibility (w/P) is 0. The rock compressibility at a
pressure of 4,500 psia is 4 x 10-6 PSI-1. Water and oil relative permeability data and
capillary pressures are given as functions of water saturation below.
Sw

krwater

kroil

0.25
0.50
0.70
0.80

0.00
0.20
0.40
0.55

0.90
0.30
0.10
0.00

capillary pres.
(psi)
4.0
0.8
0.2
0.1

This data should be inserted in the third section of the ECLIPSE data file, the
PROPS section. The form in which this data should be entered is shown in Figure
7. The following keywords are used:
PROPS
DENSITY
PVDO
PVTW
ROCK
14

Section header
Surface density of oil, water and gas phases
PVT data for dead oil relating FVF and viscosity to pressure
PVT data for water relating FVF, compressibility and viscosity
to pressure
Compressibility of the rock

Reservoir Simulation Model Set-Up

SWOF

Table relating oil and water relative permeabilities and oil-water


capillary pressure to water saturations

PROPS
Densities in lb/ft3
Oil
Water

Gas

DENSITIES
49

63

0.01 /

PVT data for dead oil


P
Bo

Vis

300
800
6000

1.0
1.1
2.0 /

PVDO
1.25
1.20
1.15

PVT data for water


P
Bw
Cw

Vis

Viscosibility

4500

0.8

0.0 /

PVTW
1.02

3e-06

Rock compressibility
P
Cr
ROCK
4500

4e-06 /

Water and oil rel perms and capillary pressure


Sw
Krw
Kro Pc
SWOF

Figure 7.
PROPS Section of input
data file.

0.25
0.5
0.7
0.8

0.0
0.2
0.4
0.55

0.9
0.3
0.1
0.0

4.0
0.8
0.2
0.1 /

Institute of Petroleum Engineering, Heriot-Watt University

15

4.6 Initial Conditions

Once the rock and fluid properties have all been defined, the initial pressure and
saturation conditions in the reservoir must be specified. This may be done in one of
three ways:
1) Enumeration
2) Equilibration
3) Restart from a previous run
1) Enumeration. In this method of initialising the model, the pressure, oil and water
saturations in each cell at time = 0 are set in much the same way as permeabilities and
porosities are set. This method is the most complicated and least commonly used. A
failure to correctly account for densities when setting the pressures in cells at different
depths will result in a system that is not initially in equilibrium.
2) Equilibration. This is the simplest and most commonly used method for
initialising a model. A pressure at a reference depth is defined in the input data,
and the model then calculates the pressures at all other depths using the previously
entered density data to account for hydrostatic head. The depths of the water-oil
and gas-oil contacts are also specified if they are within the model, and the initial
saturations can then be set depending on position relative to the contacts. (In
a water-oil system, above the oil-water contact the system is at connate water
saturation, below the contact Sw = 1.)
3) Restart from a previous run. If a model has already been run, then one of the
output time steps can be used to provide the starting fluid pressures and saturations
for a subsequent calculation. This option will typically be used where a model has
been history matched against field data to the current point in time, and various
future development scenarios are to be compared. A restart run will use the
last time step of the history-matched model as the starting point for a predictive
calculation, which may then be used to assess future performance. Time is saved
by not repeating the entire calculation.
In this example the model is being set up to predict field performance from first oil,
and thus there is no previous run to use as a starting point. The equilibration model is
to be used, with an initial pressure of 4,500 psia at 8,000 ft. The model should initially
be at connate water saturation throughout. To achieve this, the water-oil contact
should be set at 8,200 ft, 50 ft below the bottom of the model. The water saturation
in each cell will be set to the first value in the relative permeability (SWOF) table,
which is 0.25. (If any cells were located below the water-oil contact, they would
be set to the last value in the relative permeability table, which would thus have to
include relative permeability and capillary pressure values for Sw = 1.)
An output file containing initial cell pressures and saturations for display should
be requested so that a visual check can be made that the correct initial values of
these properties have been calculated.
This initialisation data should be inserted in the fourth section of the ECLIPSE
data file, the SOLUTION section. The form in which this data should be entered is
shown in Figure 8. The following keywords are used:
SOLUTION Section header
16

Reservoir Simulation Model Set-Up

EQUIL
RPTRST

Equilibration data (pressure at datum depth and contact depths)


Request output of cell pressures and saturations at t = 0

SOLUTION
Initial equilibration conditions
Datum
Pi@datum

WOC

Pc@WOC

8200

0/

EQUIL
8075

4500

Output to restart file for t=0 (.UNRST)


Restart file
Graphics
for init cond
only
Figure 8.
SOLUTION Section of
input data file.

RPTRST
BASIC=2

NORST=1 /

4.7 Output Requirements

Clearly there is no point in performing a reservoir simulation if no results


are output. The parameters that should be calculated are specified in the
SUMMARY section by the use of appropriate keywords, but for this section
only the keywords are not found in the main section of the manual, but in the
Summary Section Overview.
Most of the summary keywords consist of four letters that follow a basic
convention.
1st letter:
F - field
R - region
W - well
C - connection
B - block
2nd letter:
O - oil (stb in FIELD units)
W - water (stb in FIELD units)
G - gas (Mscf in FIELD units)
L - liquid (oil + water) (stb in FIELD units)
V - reservoir volume flows (rb in FIELD units)
T - tracer concentration
S - salt concentration
C - polymer concentration
N - solvent concentration

Institute of Petroleum Engineering, Heriot-Watt University

17

3rd letter:
P - production
I - injection
4th letter:
R - rate
T - total
Thus, use of the keyword FOPR requests that the Field Oil Production Rate be output,
and WWIT represents Well Water Injection Total, etc.
Keywords beginning with an F refer to the values calculated for the field as a whole,
and require no further identification. However, keywords beginning with another
letter must specify which region, well, connection or block they refer to. Thus,
for example, a keyword such as FOPR requires no accompanying data, but WWIT
must be followed by a list of well names, terminated with a /. If no well names
are supplied, and the keyword is followed only by a /, the value is calculated for all
wells in the model. An example would be
FOPR
WWIT
Inj /
WBHP
/
Here, the following will be calculated:
Oil production rate for the entire field
Cumulative water injection for well Inj
Well bottomhole pressure for all wells in the model
In Tutorial 1A the following parameters should be calculated and output:
Field average pressure
Bottomhole pressure of all wells
Field oil production rate
Field water production rate
Field oil production total
Field water production total
Water cut in well PROD
CPU usage
In addition, the output Run Summary file (.RSM) should be defined such that it
can easily be read into MS Excel.
The form in which this data should be entered is shown in Figure 9.
following keywords are used:
SUMMARY Section header
FPR
Field average pressure
18

The

Reservoir Simulation Model Set-Up

WBHP
FOPR
FWPR
FOPT
FWPT
WWCT
CPU
EXCEL

Well Bottomhole pressure


Field Oil Production Rate
Field Water Production Rate
Field Oil Production Total
Field Water Production Total
Well Water Cut
CPU usage
Create summary output as Excel readable Run Summary file

SUMMARY
Field average pressure
FPR
Bottomhole pressure of all wells
WBHP
/
Field oil production rate
FOPR
Field water production rate
FWPR
Field oil production total
FOPT
Field water production total
FWPT
Water cut in PROD
WWCT
PROD /
CPU usage
TCPU
Figure 9.
SUMMARY Section of
input data file.

Create Excel readable run summary file (.RSM)


EXCEL

4.8 Production Schedule

Having defined the initial conditions (t = 0) in the SOLUTION Section, the final part
of the input data file defines the well controls and time steps (t > 0) in the SCHEDULE
Section. The main functions that are performed here are:
Specify grid data to be output for display or restart purposes
Define well names, locations and types
Specify completion intervals for each well
Specify injection and production controls for each well for each given period

Institute of Petroleum Engineering, Heriot-Watt University

19

(time step)
Basic pressure and saturation data should be generated at every time step to enable
a 3-D display of the model to be viewed at each time step.
A production well, labeled PROD and belonging to group G1, is to be drilled in cell
position (1,1), and a water injection well, INJ belonging to group G2, is to be drilled
in cell (5,5). Both wells will have 8 inch diameters, and should be completed in
all three layers. They both have pressure gauges at their top perforation (8,000 ft).
Both will be open from the start of the simulation, enabling production of 10,000
stb/day of liquid (oil + water) from the system, and pressure support provided by
injection of 11,000 stb/day of water. The simulation should run for 2,000 days,
outputting data every 200 days.
A number of keywords in the SCHEDULE section will be able to read in data that
the user may wish to default or not supply all. This can be done by using the *
character, with the number of values to be defaulted or ignored on the left, and the
space to the right left blank. Thus 1* ignores one value, 2* ignores the next
two values, etc. For example, in the COMPDAT keyword, we may wish to specify
values for items 1 to 6, and item 9 (which is the wellbore diameter), but items 7 and 8
should remain unspecified. This may be achieved as follows:
Completion interval
Well
Location
name
I
J

COMPDAT
Item number
1
2
PROD
1
/

3
1

Interval
K1 K2

4
1

5
3

Status
O or S

6
OPEN

Well
ID

7 8
2*

9
0.6667 /

The keywords to be used are:


SCHEDULE Section header
RPTRST
Request output of cell pressures and saturations at all time
steps (t > 0)
WELSPECS Define location of wellhead and pressure gauge
COMPDAT Define completion intervals and wellbore diameter
WCONPROD
Production control
WCONINJ
Injection control
TSTEP
Time step sizes (for output of calculated data)
END
End of input data file
These keywords should appear as the last section of the data file as shown in Figure
10. Although not the case in this simple example, this section will typically be the
longest, containing flow rates for each well on a monthly basis for the history of the
field. It should be noted that the time steps input here refer to time intervals at which
20

Reservoir Simulation Model Set-Up

data are output. The simulator will try to use these time step sizes as numerical
time step also, but if the calculations do not converge, it will automatically cut the
numerical time step sizes.
SCHEDULE
Output to Restart file for t > 0 (.UNRST)
Restart file
Graphics
only
every step
RPTRST
BASIC=2

NORST=1 /

Location of wellhead and pressure gauge


Well
Well
Location
BHP
name group
I
J
datum
WELSPECS
PROD
INJ
/

G1
G2

1
5

Completion interval
Well
Location
name
I
J
COMPDAT
PROD
INJ
/

1
5

1
5

Production control
Status Control
Well
name
mode
WCONPROD
PROD OPEN
/

Figure 10.
SCHEDULE Section of
input data file.

WATER

8000
8000

OIL
/
WATER /

Interval
K1 K2

Status
0 or S

1
1

OPEN
OPEN

3
3

Oil
rate

LRAT

Injection control
Fluid
Status
Well
Name TYPE
WCONINJ
INJ
/

1
5

Pref.
phase

Wat
rate

Gas
rate

3*

Control
mode

OPEN RATE

Well
ID
2*
2*

Liq.
rate
10000

Surf
rate
11000

Resv
rate

0.6667 /
0.6667 /

Resv
rate
1*

BHP
lim
2000 /

Voidage
frac flag

BHP
lim

3*

20000 /

Number and size (days) of timesteps


TSTEP
10*200 /
END

Institute of Petroleum Engineering, Heriot-Watt University

21

5. RUNNING ECLIPSE AND FILE NAME CONVENTIONS


5.1 Running ECLIPSE on a PC

Once the input data file has been edited, it should be saved with the file extension
.DATA. For Tutorial 1A choose a file name such as TUT1A.DATA. Care
should be taken that the text editor does not append the suffix .txt onto the
file name, as this will render the file unreadable to ECLIPSE. This can be
avoided by using Menu->File->Save As and selecting All Files instead of Text
Documents as the Save as type.
Having saved the input data file, the GeoQuest Launcher may be used to run
ECLIPSE, and the user will be prompted to locate the input file. The simulation
will then start, and will run by reading every keyword in the order in which
they appear in the input file.

5.2 File Name Conventions

If any of the keywords or data are incorrectly entered, the run will stop without
performing the required flow calculations. If the simulator is satisfied that all data
has been entered correctly, then it will perform the requested flow calculations, and
various output files will be generated during the run, as follow:
TUT1A.PRT
The .PRT file is an ASCII file that is generated for every
successful and unsuccessful run. It contains a list of the keywords, and will
indicate if any keywords have been incorrectly entered. If the simulation fails,
this file should be checked for the cause of the failure. A search for an ERROR
in this file will usually reveal which keyword was the culprit. If the run was
successful, this file will contain summary data such as field average pressure
and water cut for each time step.
TUT1A.GRID The .GRID file is a binary file that contains the geometry of the
model, and is used by post processors for displaying the grid outline.
TUT1A.INIT
The .INIT file is a binary file that contains initial grid property
data such as permeabilities and porosities. These may be displayed using a
post-processor to check that the data have been entered correctly, and to display
a map of field permeabilities, etc.
TUT1A.UNRST The .UNRST file is a unified binary file that contains pressure and
saturation data for each time step. These may be displayed using a post-processor,
or may be used as the starting point for an ECLIPSE restart run.
TUT1A.RSM
The .RSM file is an ASCII file that can be read into MS Excel
to display summary data in line chart format. This file is only created once the run
has completed. During the run the summary data is stored in file TUT1A.USMRY,
which is a binary file readable only by the GeoQuest post-processors.
The GeoQuest post-processors are Graf and FloViz. During this course FloViz is used
for 3D displays of the model, showing, for example, progression of the water flood
22

Reservoir Simulation Model Set-Up

by displaying saturations varying with time (Figure 11). Excel is used to display line
charts such as water cut vs. time, etc. Both grid and line charts may be displayed in
Graf, which is a powerful though more complicated post-processor than FloViz. The
functionality of Graf is being replaced by ECLIPSE Office, which is a GUI that may
be used for setting up data files and viewing results, and may optionally be used for
subsequent tutorial sessions. However, students are encouraged to use a basic text
editor for pre-processing, and Excel and FloViz for post-processing for Tutorial 1A,
since this will give a better understanding of the calculations being performed.

Figure 11
FloViz visualisation of
water saturation for four
time steps, showing
progression of water flood
in Tutorial 1A. The
injection well is on the left
and the production well on
the right of the 5 x 5 x 3
model.

6. CLOSING REMARKS
In this section of the course, we have presented the working details of how to set up
a a practical reservoir simulation model. We have used the Schlumberger GeoQuest
ECLIPSE software for the specific case presented here. However, the general
procedures are very similar for most other commercially available simulators.
The various input data that are required should be quite familiar to you from the
discussion in the introductory chapter of the course (Chapter 1). However, how
these are systematically organised as input for the simulator should now be clear.
The vast possibilities for simulation output have also been discussed in this section
and you should know be aware of how to choose this output, organise it is files and
then visualise it later. The issue of visualisation was also discussed previously but
its value should be better appreciated by the student.

Institute of Petroleum Engineering, Heriot-Watt University

23

ECLIPSE TUTORIAL 1
(A 3D 2-Phase Reservoir Simulation Problem)
A.
Prepare an input data file for simulating the performance of a two-phase (water/oil)
three dimensional reservoir of size 2500 x 2500 x 150, dividing it into three layers of
equal thickness. The number of cells in the x and y directions are 5 and 5 respectively.
Other relevant data are given below, using field units throughout:
Depth of reservoir top:
Initial pressure at 8075:
Porosity:

8000 ft
4500 psia
0.20

Permeability in x direction:

200 mD for 1st and 3rd layers and 1000 mD for


2nd layer.
150 mD for 1st and 3rd layers and 800 mD for
2nd layer.
20 mD for 1st and 3rd layers and 100 mD for 2nd
layer.

Permeability in y direction:
Permeability in z direction:

500

500

500

500

500

1
2
3

50
50
50
1

500

500

500

Figure 1
Schematic of model.

500
500

Water and Oil Relative Permeability and Capillary Pressure Functions.

24

Water Saturation

krw

kro

0.25*
0.5
0.7
0.8
1.0

0.0
0.2
0.4
0.55
1.00

0.9
0.3
0.1
0.0
0.0

Pcow
(psi)
4.0
0.8
0.2
0.1
0.0

* Initial saturation throughout.

Reservoir Simulation Model Set-Up

Water PVT Data at Reservoir Pressure and Temperature.


Pressure
(psia)

Bw
(rb/stb)

cw
(psi-1)

w
(cp)

Viscosibility
(psi-1)

4500

1.02

3.0E-06

0.8

0.0

Oil PVT Data, Bubble Point Pressure (Pb) = 300 psia.


Pressure
(psia)

Bo
(rb/stb)

Viscosity
(cp)

300
800
6000

1.25
1.20
1.15

1.0
1.1
2.0

Rock compressibility at 4500 psia:


Oil density at surface conditions:
Water density at surface conditions:
Gas density at surface conditions:

4E-06 psi-1
49 lbs/cf
63 lbs/cf
0.01 lbs/cf

The oil-water contact is below the reservoir (8,200 ft), with zero capillary pressure
at the contact.
Drill a producer PROD, belonging to group G1, in Block No. (1, 1) and an injector
INJ, belonging to group G2, in Block No. (5, 5). The inside diameter of the wells is
8. Perforate both the producer and the injector in all three layers. Produce at the
gross rate of 10,000 stb liquid/day and inject 11,000 stb water/day. The producer has
a minimum bottom hole pressure limit of 2,000 psia, while the bottom hole pressure
in the injector cannot exceed 20,000 psia. Start the simulation on 1st January 2000,
and use 10 time steps of 200 days each.
Ask the program to output the following data:

Initial permeability, porosity and depth data (keyword INIT in GRID section)

Initial grid block pressures and water saturations into a RESTART file
(keyword RPTRST in SOLUTION section)

Field Average Pressure


Bottom Hole Pressure for both wells
Field Oil Production Rate
Field Water Production Rate
Total Field Oil Production
Total Field Water Production
Well Water Cut for PROD
CPU usage

(FPR)
(WBHP)
(FOPR)
(FWPR)
(FOPT)
(FWPT)
(WWCT)
(TCPU)

to a separate Excel readable file (using keyword EXCEL) in the SUMMARY


section.
Institute of Petroleum Engineering, Heriot-Watt University

25

Grid block pressures and water saturations into RESTART files at each report
step of the simulation (keyword RPTRST in SCHEDULE section)

Procedure:
1

Edit file TUT1A.DATA in folder \eclipse\tut1 by dragging it onto the


Notepad icon, fill in the necessary data, and save the file.

Activate the ECLIPSE Launcher from the Desktop or the Start menu.

Run ECLIPSE and use the TUT1A dataset.

When the simulation has finished, use Excel to open the output file
TUT1A.RSM, which will be in the \eclipse\tut1 folder. You will need to
view.
Files of type: All files (*.*)
and import the data as Fixed width columns.

Plot the BHP of both wells (WBHP) vs. time and the field average pressure
(FPR) vs. time on Figure 1.

Plot the water cut (WWCT) of the well PROD and the field oil production
rate (FOPR) vs. time on Figure 2.

Plot on Figure 3 the BHP values for the first 10 days in the range 3,500 psia
to 5,500 psia.

Explain the initial short-term rise in BHP in the injection well and drop in BHP in
the production well. Account for the subsequent trends of these two pressures and
of the field average pressure, relating these to the reservoir production and injection
rates, water cut and the PVT data of the reservoir fluids.
B.
Make a copy of the file TUT1A.DATA called TUT1B.DATA in the same folder
tut1.
By modifying the keyword TSTEP change the time steps to the following:
15*200
Modify the WCONINJ keyword to operate the injection well at a constant flowing
bottom hole pressure (BHP) of 5000 psia, instead of injecting at a constant 11,000
stb water/day (RATE).
Add field volume production rate (FVPR) to the items already listed in the SUMMARY
section.
Run Eclipse using the TUT1B.DATA file, and then plot the two following pictures
in Excel:
26

Reservoir Simulation Model Set-Up

Figure 4:
Figure 5:

Both well bottom hole pressures and field average pressure vs.
time, showing pressures in the range 3,700 psia to 5,100 psia.
Field water cut and field volume production rate vs. time.

Account for the differences between the pressure profiles in this problem and Tutorial
1A. To assist with the interpretation, calculate total mobility as a function of water
saturation for 4 or 5 saturation points, using:

MTOT (Sw ) =

K ro (Sw ) Krw (Sw )


+
o
w

and show how this would change the differential pressure across the reservoir as
the water saturation throughout the reservoir increases. From Figure 5, explain the
impact of the WWCT profile (fraction) on the FVPR (rb/day).
C.
Copy file TUT1B.DATA to TUT1C.DATA in the same folder.
This time, instead of injecting at a constant flowing bottom hole pressure of 5000 psi,
let the simulator calculate the injection rate such that the reservoir voidage created
by oil and water production is replaced by injected water. To do this, modify the
control mode for the injection well (keyword WCONINJ) from BHP to reservoir rate
(RESV), and use the voidage replacement flag (FVDG) in item 8. Set the upper limit
on the bottom hole pressure for the injection well to 20,000 psia again.
Note the definitions given in the manual for item 8 of the WCONINJ keyword. Based
on the definition for voidage replacement,
reservoir volume injection rate = item 6 + (item 7 * field voidage rate)
Therefore, to inject the same volume of liquid as has been produced, set
item 6 to 0, and
item 7 to 1.
Run Eclipse using the TUT1C.DATA file, and then run Floviz, to display the grid
cell oil saturations (these displays need NOT be printed).
Discuss the profile of the saturation front in each layer, and explain how it is affected
by gravity and the distribution of flow speeds between the wells.

Institute of Petroleum Engineering, Heriot-Watt University

27

TUT1A.DATA
RUNSPEC
TITLE
NDIVIX

NDIVIY

NDIVIZ

NCWMAX

NGMAXZ

Y1

Z1

DIMENS
OIL
WATER
FIELD
NWMAXZ
WELLDIMS
START
GRID
DX
DZ
PORO
X1
BOX
TOPS
?DBOX
PERMX
PERMY
PERMZ
INIT

28

X2

Y2

Z2

NWGMAX

Reservoir Simulation Model Set-Up

PROPS
OIL

WAT GAS

DENSITY
P

Bo

Vis

Bw

Cw

Vis

Cr

Sw

Krw

Kro

Pc

PVDO
Viscosibility

PVTW

ROCK

SWOF
SOLUTION
DATUM

Pi@DATUM

WOC

Pc@WOC

GOC

Pc@GOC

EQUIL
Block Block Create initial
P
Sw
restart file
RPTSOL
SUMMARY

RPTSMRY

Institute of Petroleum Engineering, Heriot-Watt University

29

SCHEDULE
Block
P

Block
Sw

Create restart file


at each time step

WELL
GROUP

LOCATION
I
J

RPTSCHED
WELL
NAME

BHP
DATUM

PREF.
PHASE

WELSPECS
WELL
NAME

LOCATION
I
J

INTERVAL
K1
K2

STATUS
O or S

WELL
ID

COMPDAT
WELL STATUS CONTROL OIL
NAME
MODE
RATE

WAT
RATE

GAS
RATE

LIQ
RESV
RATE RATE

BHP
LIMIT

WCONPROD
WELL FLUID STATUS
NAME TYPE
WCONINJ
DAYS
TSTEP
END

30

CONTROL
MODE

SURF
RATE

RESV
RATE

VOIDAGE
BHP
FRAC FLAG LIMIT

Gridding And Well Modelling

CONTENTS
1. INTRODUCTION
2. GRIDDING IN RESERVOIR SIMULATION
2.1. Introduction
2.2. Accuracy of Simulations and Numerical
Dispersion
2.3. Grid Orientation Effects
2.4 Local Grid Refinement (LGR)
TGrids
or (So ) i +1/ 2
(So ) iand
2.5 Distorted
1 / 2 Cornero Point
Geometry
2.6 Issues in Choosing a Reservoir Simulation

Grid

2.7 Streamline Simulationp


p

Pi Pi 1
Q p = kA. .
= kA. .

Bp x
Bp x i 1 + x i

3. THE CALCULATION OF BLOCK TO BLOCK


FLOWS IN RESERVOIR SIMULATORS
3.1. Introduction
to Averaging of Block to
kA
Block Flows
3.2. Averaging of the Two-Phase Mobility
Term, p
4. WELLS IN RESERVOIR SIMULATION
4.1. Basic Idea
Bp of a Well Model
4.2. Well Models for Single and Two-Phase
Flow
4.3 Well Modelling
in a Multi-Layer System
x
4.4. Modelling Horizontal Wells
4.5 Hierarchies of Wells and Well Controls

x i 1 + x i
2

5. CLOSING REMARKS - GRIDDING AND


WELL MODELLING

k rp
p

LEARNING OBJECTIVES:
Having worked through this chapter the student should:
understand and be able to describe the basic idea of gridding and of spatial and
temporal discretisation.
be aware of all the main types of grid in 1D, 2D and 3D used in reservoir simulation
and be able to describe examples of where it is most appropriate to use the different
grid types.
be able to give a short description with simple diagrams of the phenomena of
numerical dispersion and grid orientation and to explain how these numerical problems
can be overcome.
be familiar with more sophisticated issues in gridding such as the use of local grid
refinement (LGR), distorted, PEBI and corner point grids
given a specific task for reservoir simulation, the student should be able to select
the most appropriate grid dimension (1D, 2D, 3D) and geometry/structure (Cartesian,
r/z, corner point etc.).
be able to discuss the issues of grid fineness/coarseness (i.e. how many grid blocks
do we need to use) in terms of some examples of what can happen if an inappropriate
number of grid blocks are used in a reservoir simulation calculation.
be able to describe the basic ideas behind streamline simulation and to compare it with
conventional reservoir simulation in terms of its advantages and disadvantages.
be familiar with the different types of average used for single phase kA , two phase
relative permeabilities (krp) and for p and Bp (p = o, w, g phase) when calculating
the block to block flows (Qp) in a reservoir simulator.
be able to describe the physical justification for using the upstream value of two
phase relative permeabilities when calculating the block to block flows (Qp) in a
reservoir simulator.
understand the origin of all the pressure drops that are experienced by the reservoir
fluids from deep in the reservoir, through to the wellbore and then to the surface
facilities and beyond
know what a well model is and what productivity index (PI) is, including knowing
the radial Darcy Law and how this gives a mathematical expression for PI for single
phase flow (know the expression from memory).
be able to describe the main issues in relating the pressure in the reservoir, Pe, at
some drainage radius, re, to an average grid block pressure and how this leads to the
Peaceman formula (r = 0.2 x) which is then used to calculate PI.
be able to extend PI to the concept of multi-phase flow to calculate PIw and PIo.
2

Gridding And Well Modelling

describe a well model for a multi layer system where there is two phase flow into
the wellbore.
understand and be able to describe the various types of well constraint that can be
applied e.g. injection volume constrained wells, well flowing pressure constraints
and voidage replacement constraints.

GRIDDING AND WELL MODELLING IN RESERVOIR


SIMULATION
This chapter deals with the related issues of grid selection and well modelling in
reservoir simulation.
Gridding: Examples of various grid geometries are presented including 1D linear,
2D areal, 2D cross-sectional , 2D radial (r/z) and 3D Cartesian cases. Selection
of grid geometry, how fine a grid to take and potential problems are discussed.
Non-mathematical introductions to the concepts of numerical dispersion and grid
orientation are given along with some examples of the consequences of these
effects. More sophisticated gridding such as PEBI grids, distorted grids and local
grid refinement (LGR) are illustrated with examples. A brief discussion of streamline
simulation is also presented. Treatment of the block to block flows in reservoir
simulation is presented and it shown how the various inter-block quantities such as
the permeability and relative permeabilities are averaged.
Well Modelling: All interactions between the surface facilities and the reservoir
takes place through the injector and producer wells. It is therefore very important
to model wells accurately in the reservoir simulation model. The central issue with
a well model in a simulator is that it must represent a near singular line source within
(usually) a very large grid block. The basic ideas of well modelling are explained and
how simple well controls are applied is introduced. Modelling of horizontal wells
and the control of well hierarchies are also briefly discussed.

GRIDDING IN RESERVOIR SIMULATION

2.1 Introduction

By this point in the course, you will be familiar with the idea of gridding since it
has been discussed in Chapter 1 both in general terms and with reference to the
SPE examples (Cases 1 - 3; Section 1.3). You will also have seen how to set up a
3D grid in Chapter 3. Basically, the gridding process is simply one of chopping the
reservoir into a (large) number of smaller spatial blocks which then comprise the
units on which the numerical block to block flow calculations are performed. More
formally, this process of dividing up the reservoir into such blocks is known as spatial
discretisation. Recall that we also divide up time into discrete steps (denoted t) and
this related process is known as temporal discretisation. The numerical details of
how the discretisation process is carried out using finite difference approximations
of the governing flow equations is presented in Chapter 6. However, we would point
out that the grid used for a given application is a user choice - it is certainly not a
Institute of Petroleum Engineering, Heriot-Watt University

reservoir given - although, as we will see, there are some practical guidelines that
help us to make sensible choices in grid definition.
In a Cartesian grid, which to date has been the mot common type of grid used, we
denote the block size as x, y and z and these may or may not be equal. This grid
can also be in one dimension (1D), two dimensions (2D) or three dimensions (3D).
Some typical examples of 1D, 2D and 3D Cartesian grids are shown in Figure 1.
This is the most straightforward type of grid to set up and typical application of such
grids are as follows:
- 1D linear grids may be used to simulate 1D Buckley-Leverett type water displacement
calculations (x-direction) or for single column vertical displacements (z-direction)
such as gravity stable gas displacement of oil (Figures 1(a) and 1(b));
- 2D Cartesian grids: 2D cross-sectional (x/z) grids may be used to; (a) study
vertical sweep efficiency in a heterogeneous layered system; (b) calculate water/oil
displacements in a geostatistically generated cross-section; (c) generate pseudo-relative
permeabilites (can be used to collapse a 3D calculation down to a 2D system); (d) to
study the mechanism of a gas displacement process - e.g. to determine the importance
of gravity etc. (Figure 1(e));
2D areal (x/y) grids may be used to; (a) calculate areal sweep efficiencies in a
waterflood or a gas flood; (b) to examine the stability of a near-miscible gas injection
within a heterogeneous reservoir layer; (c) examine the benefits of infill drilling in
an areal pattern flood etc. (Figures 1(c) and 1(d));
- 3D (x/y/z) Cartesian grids are used to model a very wide range of field wide reservoir
production processes and would often be the default type of calculation for a typical
full field simulation of waterflooding, gas flooding, etc. (Figure 1(f)).
Cartesian grids are clearly quite versatile but they are not appropriate for all flow
geometries that can occur in a reservoir. For example, close to the wellbore (of a
vertical well), the flows are more radial in their geometry. For such systems, an r/z
- geometry may be more appropriate as shown in Figure 2. An r/z grid is frequently
used when modelling coning either of water or gas into a producer. The pressure
gradients near the well are very steep and, indeed, we know from the discussion in
Chapter 2, section 3.5 that the pressure varies as ln(r/rw), where rw is the well radius.
For this reason, in coning studies, a logarithmically spaced grid is often used for
the grid block size, r; a logarithmic spacing divides up the grid such that (ri/ ri-1) is
constant where ri = (ri- ri-1) as shown in Figure 2. For example, if we take rw = 0.5
ft and the first grid block size is, r1 = 1ft, then this sets (ri/ ri-1) = 3 since r0 = rw =
0.5ft and r1 = 1.5ft.; hence r2 = 3x1.5ft. = 4.5ft. and r2 = 3ft., r3 = 13.5ft and r3
= 9ft., and so on.

Gridding And Well Modelling

(a) 1D horizontal grid

(c) 2D Areal grid - top view showing


injectors ( ) and producers ( )

(b) 1D vertical grid


1D vertical displacement
e.g. 1D vertical gravity
drainage calculations
z

W2

(d) 2D areal grid

W3

W1
y
x

x
y

Perspective view of
a 2D areal (x/y)
reservoir simulation
grid: W = well

Just 1 x z - block in 2D areal grid

(e) 2D (x/z) cross-sectional model showing a waterflood

Institute of Petroleum Engineering, Heriot-Watt University

Water Injector

Producer

(f) a 3D Cartesian grid with variable vertical grid (z varies from layer to layer)

Water Injector

Producer

Figure 1
Examples of 1D, 2D and 3D
Cartesian grids

y
z
x

top view

ri
rw

z
h
r

Notation:

r
r
z
zi
h
rw

zi

ri
ri-1

= radial distance from well


= radial grid size (can vary r1, r2, ...)
= vertical coordinate
= vertical grid size (can vary z1, z2, ...)
= height of formation
= wellbore radius

2.2 Accuracy of Simulations and Numerical Dispersion

The issue of numerical dispersion was touched upon briefly in Chapter 1 in connection
with Case 1 (SPE10022). Here we expand on the concept in a non-mathematical
manner. Numerical dispersion is essentially an error due to the fact that we use
a grid block approximation for solving the flow equations. A more mathematical
description of numerical dispersion is presented in Chapter 6, Section 8 but here we
focus on explaining physically how it arises and what its consequences are. We also

Figure 2
r/z grid geometry - more

appropriate for modelling


flows in the near well
region even in a
heterogeneous layered
system as shown.

Gridding And Well Modelling

discuss how to reduce this source of error in our simulations or to make it a minor
effect. The balance, as we will see, is between accuracy (usually by taking more
grid blocks) and computational cost. Ideally, we would like to capture all the main
reservoir processes (e.g. frontal displacement, crossflow, gravity segregation etc.) and
accurately forecast recovery to some acceptable percentage error, for the minimum
number of grid blocks.
A simple schematic of the way that the numerical dispersion error arises is shown
step by step in Figure 3. This figure illustrates a simple linear waterflood and we
imagine that each of the sub-figures shown from (a) to (e) represents a time step, t,
of the water injection process. Block i = 1 contains the injector well which is injecting
water at a constant volumetric rate of Qw, and block i = 5 contains the producer. The
system is initially at water saturation, Swc, and this water is immobile i.e. the relative
permeability of water is zero, krw(Swc) = 0. Each block has constant pore volume, Vp
= x.A. where is the porosity. In Figure 3(a), we see that after time, t = t, some
quantity of fluid has been injected into block i = 1; the volume of water injected is
Qw. t and this would cause a water saturation change in grid block i = 1, Sw1 = (Qw.
t)/Vp i.e. the new water saturation in this block is now, Swc+ (Qw. t)/Vp. Over the
first time step, no fluid flowed from block to block since the relative permeability of
all blocks was zero (krw(Swc) = 0). However, the relative permeability in block i = 1
is now krw(Sw1) > 0. The second time period of water injection is shown in Figure
3(b). Another increment of water, Qw. t, is injected into block i = 1 causing a further
increase in water saturation Sw1. However, over the second time period, krw(Sw1) > 0
and therefore water can flow from block i = 1 to i = 2, increasing the water saturation
such that Sw2 > Swc making the relative permeability in this block, krw(Sw2) > 0. In
the third time step, shown in Figure 3 (c), the same sequence occurs except that fluid
can now from block 1 2 and also 2 3, where for the same reasons as explained,
krw(Sw3) > 0. In the fourth and fifth time steps (Figures 3(d) and 3(e)), flow can
now go from block 3 4 and from 4 5 where, since krw(Sw5) > 0, then it can be
produced from this block, although the relative permeability in block 5 will be very
small. Hence, after only five time steps to time , t = 5t, the water has reached the
producer in block 5 from whence it can be produced (although not a very high rate
because the relative permeability is very small) and this is an unsatisfactory situation.
If we had taken 10 grid blocks, then clearly a similar argument would apply and water
would be produced after 10 time steps - with an even lower relative permeability in
block i = 10 - and this is more satisfactory. Indeed, this underlies why we take more
grid blocks. This simple illustration explains in a quite physical way the basic idea
of numerical dispersion.

Institute of Petroleum Engineering, Heriot-Watt University

1.0

Relative Permeability to Water

krw
Swc = 20%

20

Water
Injection

Qw

Sor = 30%

40
60
Sw %

80

100
Oil
Production

Mixing due to numerical dispersion


& relative permeability effects

Time Step
(a)

Sw1

i=1

i=2

i=3

Sw = Swc
i=4

1
t = t

i=5

(b)

2
t = 2.t

(c)

3
t = 3.t

(d)

4
t = 4.t

(e)

5
t = 5.t

The frontal spreading effect of numerical dispersion can be seen when we try to simulate
the actual saturation profile, Sw(x,t), in a 1D waterflood. Under certain conditions,
this may have an analytical solution, e.g. the well known Buckley-Leverett solution
described in Chapter 2, Section 4.2. This often has a shock front solution for the
advancing water saturation profile, Sw(x,t) as shown in Figure 4. Clearly, this sharp
front may be lost in a grid calculation since, at time t, the front will have a definite
position, x(t). However, in a grid system with block size x, any saturation front
can only be located within x as shown in Figure 4. If we take more grid blocks
(x decreases), then we will locate the front more accurately. Indeed, taking more
and more blocks we will gradually get closer to the analytical (correct) solution.
Hence, one method of reducing numerical dispersion is to increase the number of
grid blocks. An alternative is to use a numerical method which has inherently less
dispersion in it but we will not pursue this here. Another approach is to use pseudo
functions to control numerical dispersion - as we briefly introduced in Chapter 1
- and this is discussed further below.

Figure 3
Effect of grid on water
breakthrough time - numerical dispersion.

Gridding And Well Modelling

1.0
Numerical Grid block size = x
Water saturation = Sw(x,t1)

time = t1

Front moving in this direction


with velocity Vw
Analytical
solution

Sw

Figure 4
The frontal spreading of
a Buckley-Leverett shock
front when calculated using
a 1D grid block model

xf(t1)

2.3 Grid Orientation Effects

Figure 5
Flow between an injector
(I) and 2 producers (P1
and P2) where the injectorproducer separations
are identical but flow is
either oriented with the
grid or diagonally across
it illustrating the grid
orientation effect.

Another numerical problem arising in 2D and 3D grids is the grid orientation effect.
This is illustrated in Figure 5 where the distance between wells I - P1 and I - P2 are
the same. However I - P1 are joined by a row of cells oriented to the flow as shown.
The flow between I - P2 is rather more tortuous as also shown in Figure 5. The Grid
Orientation effect arises when we have fluid flow both oriented with the principal
grid direction and diagonally across this grid. Numerical results are different for
each of the fluid paths through the grid structure. This problem arises mainly due
to the use of 5-point difference schemes (in 2D) in the Spatial Discretisation. It
may be alleviated by using more sophisticated numerical schemes such as 9-point
schemes (in 2D).

P1
I

I = Injector
P = Producer

P2

Flow arrows show


the fluid paths in
oriented grid and
diagonal flow
leading to grid
orientation errors

The effects on the breakthrough time and in recoveries of these two flow orientations
are shown in Figure 6. The I-P1 orientation tends to lead to somewhat earlier
breakthrough and a less optimistic recovery that the I-P2 orientation. The reasons
for this are intuitively fairly obvious.

Institute of Petroleum Engineering, Heriot-Watt University

diagonal flow result


"true"
recovery

Oil
recovery

Time

grid
aligned
result

Figure 6
Oil recoveries for true,
aligned and diagonal flows
in 2D grid

The grid orientation error can be emphasised for certain types of displacement. For
example, in gas injection, gas viscosity is much less than the oil viscosity (g << o)
leading to viscous fingering instability. This is exaggerated when flow is along the
grid as in the I-P1 orientations.
Again, as for numerical dispersion, some grid refinement can help to reduce the grid
orientation effect. Alternative numerical schemes can also be devised to reduce this
source of error. In particular, in a 2D grid the flows are usually worked out using the
neighbouring grid blocks shown in Figure 7. The 5-point numerical scheme uses the
neighbours shown in Figure 7(a). If information from the blocks in Figure 7(b) are
used, the 9-point scheme that emerges helps greatly to reduce the grid orientation
error. We will not go into further technical details here.
(a)

(b)

Figure 7
5 - point and 9 - point
schemes for discretising the
grid - the latter helps to
reduce grid orientation
effects

2.4 Local Grid Refinement (LGR)

In a reservoir, the changes in pressure, saturations and flows tend to be quite different
in different parts of the system. For example, close to a well which is changing
production rate every day or so, there will be large pressure and saturation changes.
On the contrary, on a flank of the field which is connected to, but is remote from,
the active wells, the pressure may be quite slowly changing and the saturations may
hardly be changing at all. To represent regions with rapidly changing waterfronts will
require a finer grid than will be required for relatively stagnant regions of the system.
Thus, a single uniform grid with fixed x, y and z will often not be suitable to
represent all regions of an active reservoir. Instead, the application of some local
grid refinement (LGR) may be much more appropriate.
LGR options are supported by most major simulation models and the simplest version
is shown in Figure 8 (indeed, this simpler version is sometimes not referred to as
LGR). True LGR is shown in Figure 9 where the refined grid is clearly seen.

10

Gridding And Well Modelling

Figure 8
A simple version of local
grid refinement where the
grid is finer in the (central)
area of interest in the
reservoir

Figure 9
True local grid refinement (LGR) where the
refined grid is embedded in
the coarser grid.

In addition to conventional LGR (Figure 9), we can also define Hybrid Grid LGR
as shown in Figure 10. Hybrid Grids are mixed geometry combinations of grids
which are used to improve the modelling of flows in different regions. The
most common use of hybrid grids are Cartesian/Radial combinations where
the radial grid is used near a well. Hybrid Grid LGR can be used in a similar
way to other LGR scheme.
Schematic of Local Grid Refinement (LGR)

injector
producer

Figure 10
A simple example of LGR
and Hybrid Grid structure

Coarse grid
in aquifer

Hybrid Grid

2.5 Distorted Grids and Corner Point Geometry

In recent years, many studies have used grids that have tried to distort either to
reservoir geometry or to the particular flow field and an examples of a distorted
grid is shown in Figure 11.

Institute of Petroleum Engineering, Heriot-Watt University

11

Figure 11
An example of a distorted
grid

An interesting class of non-Cartesian grids are PEBI grids; PEBI = Perpendicular


Bisector. In the PEBI grid the chosen grid points are blocked off into volumes using
a geometrical construction which is not shown here. PEBI grids have been developed
very extensively by Aziz and co-workers at Stanford University and by Heinemann
in Austria (Heinemann, et al 1991; Palagi and Aziz, 1994).
An example of a study using PEBI grids is shown in Figure 12 where the particularly
flexible form of this grid is used to model faults in this particular case.

PEBI grids can be orientated


to follow major reservoir
faults
This example from:

R.E Phelps, T. Pham and A.M. Shahri,


Rigorous Inclusion of Faults and Fractures
in 3D Simulation, SPE59417, 2000 SPE
Asia Pacific Conference, Yokohama, Japan,
25-26 April 2000

Figure 12
An example of a study using
a PEBI grid

Another way of building distorted grids where the individual blocks retain some
broad relationship with an underlying Cartesian form is using corner point geometry
(Ponting, 1992). This is shown in Figure 13. This scheme is implemented in the
reservoir simulator Eclipse (GeoQuest, Schlumberger) where it has been applied
quite widely. In corner point geometry it appears rather tedious to build up a grid by
specifying all 8 corners of every block (although some are shared with neighbours).
However, if this approach is used, the engineer would virtually always have access to
grid building software although building complex grids can still be time consuming.
The engineer may be reluctant to use corner point geometry if there is a high likelihood
12

Gridding And Well Modelling

that the reservoir model will change radically. In future, this may be overcome by
auto-generating the corner point mesh directly from the geo-model (although some
upscaling may also be necessary in this process).
At present, corner point geometry is probably more common in fairly fixed base case
reservoir models that the engineer has fairly high confidence in. The model shown
in Figure 14 is constructed using corner point geometry since it has a major fault in
it between the aquifer and the main reservoir.
Corner Point Geometry

Coordinates of
vertices ( )
specified.
Block centres ( )

Figure 13
Corner point geometry

Highly distorted
grid blocks

Block <-> Block


Transmissibility

Figure 14
Complex reservoir model
constructed using corner
point geometry

2.6 Issues in Choosing a Reservoir Simulation Grid

The main issues in choosing a grid for a given reservoir simulation calculation are
as follows:
(i) Grid Dimension: Refers to whether we should use a 1D, 2D or 3D grid
structure;

Institute of Petroleum Engineering, Heriot-Watt University

13

(ii) Grid Geometry/Structure: The next issue is whether we should use a simple
Cartesian grid (x, y, z) or some other grid structure such as r/z. This choice also
includes where local grid refinement, a distorted grid or corner point geometry is
appropriate;
(iii) Grid Fineness/Coarseness: How many grid blocks do we need to use? This
asks whether a few hundred or thousand is adequate or whether we need 10s or 100s
of thousands for an adequate simulation calculation.
We remind the student of the advice in Chapter 1. This was to carry out the reservoir
simulation calculation keeping firmly in mind the question which had to be answered or
the decision which had to be taken. Therefore, the issues of grid dimension, type and
fineness are directly related to the appropriate question/decision. However, as we will
see, there are several technical considerations that can guide us in these choices.
Essentially, all 3 choices (grid dimension, type and fineness) depend strongly on
the problem we are trying to solve. Consider the issues of grid dimension and type
together. A 2D x/z cross-sectional model (with dip if necessary) may be used to study
the effects of vertical heterogeneity - layering for example - on the sweep efficiency
or water breakthrough time. For a near-well coning study, an r/z grid is usually more
appropriate since it more closely resembles the geometry of the near well radial flow.
2D x/z grids are also used to generate pseudo relative permeabilities for possible use in
2D areal models. For full field simulations, 3D grids are generally used which in most
models are still probably Cartesian with varying grid spacing in all three dimensions.
In recent years, other types of grid such as distorted or corner point grids are being
applied - especially if a geocellular model has been generated as part of the reservoir
description process. Such guides are also applied in some studies to model major
faults in reservoirs. Flow through major faults can lead to communication between
non neighbour blocks and this can be modelled in some simulators by defining nonneighbour grid block connections as shown schematically in Figure 15.

Fault
L1
L2

L2

L3

L1
L4

L3
L4

The issue of grid fineness/coarseness, or how many grid blocks to use in a given
simulation, can sometimes be quite subtle as we will show below. However, in
many practical calculations, some reasonable and practical number of grid block is
chosen by the engineer. Then, this can be checked by refining the grid and seeing if
the answers are close enough to the coarser calculation. If, as we carry out this grid

14

Figure 15
Grid system at a fault which
may have non-neighbour
connections

Gridding And Well Modelling

refinement, the answer - e.g. recovery profiles and water etc. - no longer changes, the
calculation is said to be converged and is probably quite reliable. By reliable,
we mean that the grid errors are probably a small contribution of to the overall
uncertainties of the whole calculation. If the grid is far from being converged, then
comparisons between different sensitivity calculations may be masked by numerical
errors. These various points are illustrated by the two examples below.
The two examples used to show the importance of the number of grid blocks are:
(i) Example 1: the effect of vertical grid fineness (i.e. NZ blocks) on a miscible
water-alternating-gas (MWAG) process.
(ii) Example 2: resolving the vertical equilibrium (VE) limit of a gas
displacement calculation.
Example 1: Figures 16(a) and 16(b) shows the recovery results (at a given time
or pore volume throughput) for both a waterflood and an MWAG flood in the same
system each as a function of 1/NZ. The difference between the two calculations is
the incremental oil recovered by the MWAG process. The economics of performing
MWAG depends on how large this difference is. The purpose of plotting this vs.
(1/NZ) is that we can extrapolate this to zero i.e., effectively to NZ . Taking
the results at NZ = 2 (1/NZ = 0.5) shows an incremental recovery of (72% - 36%)
= 36% of STOIIP which is a huge increase and would make such a project very
attractive. However, as we refine the vertical grid, the waterflood recovery increases
while the MWAG recovery decreases, i.e. the calculations move closer together and
the incremental oil is greatly reduced. Indeed, as we extrapolate to (1/NZ) = 0, we
see that the incremental oil is only (47.5% - 47%) = 0.5% which is well within the
error band of the calculation. So, rather than having a very attractive project, we
appear to have a completely marginal or non-existent improved oil recovery scheme.
Certainly, performing just one coarse grid calculation and taking the results at face
value would be very misleading in this case.
Example 2: The vertical equilibrium (VE) condition in a gas flood is where the
gas if fully segregated by gravity from the oil. This limit has a simple analytical
form (not discussed here) which can be written down without doing a grid block
calculation. However, we can test the numerical simulation by seeing how many
blocks (NZ again) we need to correctly reproduce the VE limit. The answer is rather
surprising as shown by the results in Figure 17. These results show that 200 layers
are needed to fully resolve the gas tongue at the top of the reservoir. Clearly, if
we just guessed that 5 vertical blocks would be enough and did not check, then our
calculation would be significantly in error.
The two examples above illustrate how the number of blocks chosen for a simulation
can strongly affect the results. It shows the need to check that a calculation has
converged or that changing the number of grid blocks does not significantly change
the answers.

Institute of Petroleum Engineering, Heriot-Watt University

15

Figure 16: (a) Extrapolation of Predicted Waterflood Recovery Efficiency for 2D


Stratified Model C Sand Base Case
100

Process ==> Waterflooding

90

Vertical Grid Refinement


NX

Recovery Efficiency,
Recovery
%ooIPEfficiency, %ooIP

80
70
100

Process ==> Waterflooding

60
90

Vertical Grid Refinement


As vertical gridNX
is refined 1 --> 0
NZ

50
80
40
70
30
60
Extrapolated RE = 27.6%

20
50

Water

1
2
3

Oil
NZ

1
2
3

Gas

Oil

Homogeneous Model, kv/kh = 0.1


Stratified Model, 1.2 HCPVI,
0.02
NZkv/kh = Gas
Homogeneous
1 Model, kv/kh = 0.01

As vertical grid is refined

NZ

--> 0

10
40

2D cross-sectional model

0
30
0.00 0.05

0.10

0.15

0.20 0.25 0.30


1

Homogeneous
= 0.1 0.55
0.35 0.40 Model,
0.45kv/kh
0.50
Stratified Model, 1.2 HCPVI, kv/kh = 0.02

0.60

/nz grid blocksHomogeneous Model, kv/kh = 0.01

Extrapolated RE = 27.6%

20

Recovery Efficiency,
Recovery
%ooIPEfficiency, %ooIP

10
Vertical Grid Refinement
100
NX
0
(b) Extrapolation
of Predicted
MWAG Recovery Efficiency for 2D Stratified
1
0.00 0.05 0.10 0.15 0.20 0.25 0.30
0.35 0.40 0.45 0.50 0.55 0.60
Water
90 Case
2
1
C Sand Base
/nz grid blocks
3
Oil
80
70
100
60
90
50
80
40
70
30
60
20
50

NZ
Vertical Grid Refinement
As vertical grid isNX
refined 1 --> 0
NZ
1
2
3

As vertical
is refined
Extrapolated
REgrid
= 35.3%

1 -->
NZ

Gas
Water
Oil

Homogeneous
Model, kv/kh = 0.1
NZ
Gas
Model, side solver
Variable Width Homogeneous
0Stratified Model, 25% slug, 1.2 HCPVI, kv/kh = 0.02
Homogeneous Model, kv/kh = 0.01

Process ==> Wiscible Water - Alternating - Gas (MWAG)

10
40
Extrapolated RE = 35.3%

0
30
0.00 0.05
20

0.10

0.15

0.20

Homogeneous Model, kv/kh = 0.1


Variable Width Homogeneous Model, side solver
Stratified Model, 25% slug, 1.2 HCPVI, kv/kh = 0.02
kv/kh 0.45
= 0.01 0.50 0.55 0.60
0.25Homogeneous
0.30 0.35Model,
0.40

grid blocks
Process ==> Wiscible/nzWater
- Alternating - Gas (MWAG)

10
0

0.00 0.05

16

Figure 16
The effect of vertical grid
refinement on recovery in
(a) a waterflood and (b) a
MWAG displacement in a

Water

0.10

0.15

0.20 0.25 0.30


1

0.35

/nz grid blocks

0.40

0.45 0.50 0.55 0.60

Model

Figure 17
Resolving the gas tongue
in the Vertical Equilibrium
(VE) limit in a gas - oil
displacement by increasing
the number of vertical grid
blocks (from Darman et al,
1999)

Recovery factor

Gridding And Well Modelling

0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0 x

Gravity Dominated

x x

x
x
x

0.2

0.4

Gas

x x
x x
x x
x x
x
x x

Oil

VE Limit

x x

fine grid : 5 layers


fine grid : 25 layers
fine grid : 50 layers
fine grid : 200 layers
coarse grid

PVI

0.6

0.8

2.7 Streamline Simulation

The problem of numerical dispersion was discussed above. One approach to have
a more accurate transport calculation is to use streamlines. We will describe this
qualitatively in a non-mathematical manner with reference to Figure 18 from the work
of Gautier et al (1999). The basic procedure in streamline simulation for a given
permeability field (Figure 18(a)) is to calculate the pressure distribution by solving
a conventional pressure equation (see Chapter 5 and 6). From this the iso-potentials
(pressure contours) can be calculated as shown in Figure 18(b); the gradient of the
pressures locally perpendicular to the iso-potentials are the streamlines as shown in
Figure 18(c). These streamlines are essentially the paths of the injected fluid from
the injectors (sources) to the producers (sinks). Since the velocity along these paths
is known (from Darcys Law using the calculated P), we can work out how far the
saturation front moves along the streamline, l = v.t, where v is the (local) velocity
at that point on the streamline. Since v is known quite accurately, the advance of the
front along the streamline can be calculated accurately without the problem of block
to block numerical dispersion. After we propagate the front along the streamlines, the
saturations will change over the reservoir domain. These saturation changes are then
projected back onto the Cartesian grid as shown in Figure 18(d), hence changing fluid
mobilities. These updated mobilities can be used to recalculate the pressures which,
in turn, can be used to update the streamline pattern. This process can be continued
throughout the calculation. However, the calculation of the pressure equation is what
takes most computational time in most reservoir simulation.

Institute of Petroleum Engineering, Heriot-Watt University

17

Outline of Streamline Method


Prod
a) Permeability map with an
injector and a producer wells

Inj

b) Solve the pressure field


c)

Compute the velocity field


and trace streamlines

(b)

(a)

d) Move saturation along


streamlines and compute the
values of the saturation on
the grid.

(d)

(c)

From Gautier et al (1999)

Figure 18
Schematic of streamline
simulation from the work of
Gautier et al (1999)

2D Areal Displacements with Streamlines


I1
Pair Injector / Producer

P1

IW flows faster in a
direct line between the
wells and slower in
the corners
Arrival of different
streamlines at producer
at the same time.

Injected water concentration

An example of a streamline calculation in a five-spot pattern is shown in Figure


19. In streamline simulation, it may be possible to recalculate the pressures after
many transport (saturation update) time steps. This relies on the assumption that the
streamlines do not vary too rapidly as the flood progresses. This is a good assumption
for many applications. Clearly, if the wells change very significantly, or we switch
off some wells and add new ones, it will almost certainly be necessary to recalculate
the pattern of streamlines in the reservoir domain.
Streamline simulation has gained some popularity in recent years since 3D streamline
codes have been developed e.g. by (Blunt and coworkers at Imperial College in
London) and are available commercially.
Streamline simulation is fastest compared with conventional simulation for viscous
dominated flow where the assumption of slowly changing streamlines is probably
best. For flow where gravity effects are very prominent, there tends to be side flow
between streamlines and hence it is necessary to recalculate the pressure field quite
often. This slows the streamline simulation down quite significantly in many cases
although it can sometimes remain competitive with conventional simulation. At the
present time, streamline simulation has a place in our simulation toolbox but it
18

Figure 19
An example of a streamline
calculation in a five-spot
pattern

Gridding And Well Modelling

is not suitable for all types of reservoir calculation. It is strictly for incompressible
flow and compressibility effects can cause some errors. Also, streamline codes tend
not to be as developed in terms of bells and whistles as conventional simulators
e.g. complex well models, difficult PVT oil behaviour etc.

THE CALCULATION OF BLOCK TO BLOCK FLOWS IN


RESERVOIR SIMULATORS

3.1 Introduction to Averaging of Block to Block Flows

In reservoir simulation, block to block flow terms arise between blocks which we often
denoted by mobility terms such as (T(S0))i-1/2 or (0(S0))i+1/2 etc, where the (i+1/2)
and (i-1/2) denote the block boundary as the location where that term is evaluated.
However, we do not specify properties directly on the boundaries, instead we define
them within the grid blocks themselves e.g. the saturations (Sw and So), the permeability,
porosity, relative permeability etc. A typical block-to-block flow is shown in Figure
20 where the appropriate terms in Darcys Law are also shown:

Applying
Darcys
law to the inter block flow shown in Figure 20 and using the
T (So ) i 1 / 2 or o (So ) i +1 / 2
notation in that figure, we obtain:

( (S ))
T

or ( o (So ))i +1(/ S


( T 2 o ))i 1/ 2

( T (So ))i 1/ 2 or ( o (So )

i 1 / 2

p P
p Pi Pi 1
Q p = kA. . ( =(S kA
)) . 2 por .(o(Sxoi )1)i++1/2 x i
Bp x T o i 1/ B


. p . P = kA. p
Q p2 = kA

B (1)
PB.
x . Q p p= . kA
Q pp = kA
p

Bp x
where thekA
overbar denotes an average of that
pquantity.
P The issue
phere
is: Pwhich

i Pi 1
Q
kA
=

.
.
kA
.
.
=

average should we take? The twopmain specific


are:

B questions

p x
Bp x i 1 + x i
kA

kA
2
What is the correct average for the permeability-area
product,
?

S
(
)
(
)
T
o i 1 / 2 or ( o (So ))kA
p
i +1 / 2
What is the correct average for
kAthe phase mobility, p ?

Further issues involve (i) whether we should average the


terms
k
and

p separate

P
Pi P
rp
Q p = kA. Bp . = kA. p .

p - within the phase mobility, p ; and (ii) which average


use for

Bp B
Bp wexshould
p x i 1 +
Bp from
x formation volume factor does not usually vary
although the
very rapidly
block
2
Bp

to block.

x
Bp
x
x i 1 + x i
x
3.2 Averaging
of the k-A Product, kA
=
We first examine 2the kA averaging by considering what
this
be for singlex i should
1 + x i
x + x

x
= and the related
phase flow. Consider the volumetric flow, Q, of a single phase
x i i 1
x i pressure
1 + =
2
=
drops as shown
2
p
k rp in Figure 21:
2
x i 1 + x i
p
k rp
=
k rp
2 Bp
k
rp
p
p
p
k rp
x
p
Institute of Petroleum Engineering, Heriot-Watt University

x i 1 + x i
2

19

( (S ))
T

( (S ))
Permeability = ki-1

Qp

kA

Area, A i-1

Notation:
Qp

i 1 / 2

or ( o (So ))i +1/ 2

or ( o (So ))i +1/ 2

p P
p Pi Pi 1
Q p = kA. .
= kA. .

Bp x
Bp x i 1 + x i
x i

p Pi Pi 1 2
Permeability =kpi P
Q p = kA. .
= kA. .

Bp x
Bp x i 1 + x i

2
kA
T

x i-1

i 1 / 2

Area, A i

Figure 20
Block to block flow in a
simulator

p
Bp

= volumetric flow of phase p (p = o, w, g)

Bp

xi-1, xi

= sizes of the (i-1) and i blocks (may not be equal)

= distance between grid block centres =

Ai-1, Ai

= areas of the (i-1) and i blocks


x (may
+ xnot be equal)

i 1

x i 1 + x i
2

k rp
2
= permeabilites of the (i-1) and i blocks (usually not equal)
p
k
= mobility of phase p = rp
p

ki-1, ki
p
krp

= relative permeability of phase p

= pressure drop between grid block centres

p, Bp

= viscosity and formation volume factors of phase p

Ai

Ai-1

Pi-1

Pi

ki-1

ki

xi-1

xi

Pi-1

Pressure
P

Pf = face pressure
between blocks

Pf
Pi

We now consider flows from Block (i-1) to the interface of the two blocks, where we
denote the interface pressure as Pf (see Figure 21).

20

Figure 21
Single-phase flow between
blocks to determine the
correct kA average to use in
flow simulation.

Gridding And Well Modelling

k .A
P P
x i 1
.

k
A
P
P
Q = i 1 i 1. f
1x
i 1
Q=
. x i 1i

x i 1 2

2
k .A 1 Pf 2Pi 1
2
from which
difference (Pf - Pi-1):
Q =wecani 1findi the
. pressure
x i 1

k .A
P P 1 x 1
Q = ( Pi 1 Pi 1 .) f= 2ixQ
.x i 1i 11 1

x.
i .
1
iQ
1x
k.A )i 1
((PPf (PPPif1f)) P==i 1i)1 =Q
i 1 2 2 1
) i 1
Q.2 2 ( k.A )(k(.A
f
i 1
2 ( k.A )ii 11
x
1
(Pf Pi 1 ) = Q. i 1
2 Pf ( k.A )i 1
k.Ai .A iP Pi
k
P
i
i
i
= i Pi .Pf. x fi 1 1
In block (i):
QkkQ
=i ..A
A
. P P . xi x i
=

P
P
i
(Q
)

Q f= i 1 i=. ixQi f
x i
22 ( k.A )i 1

2
k .A i Pi 2 Pf
Q =wecani find
. the2 pressure difference (P - P ):
from which
x i
i
f

k i .A i Pi 2 Pf

Q = ( P P. ) = xQ.xi1xi 1 1
xi. Qxi . 1
((PPi (PPPfi))i ==Pf )f =Q
k.)A)i
i 2 2 (k(.A

.
i
i
f
2 22 (( kk..A
A ))i
x i 1 i
(Pi Pf ) = Q.

2 Q
(k .A
)xix i 1 xx i

xtoeliminate
Adding equations( P3i and
5 above
i gives that:
xP+ term
x .1 i 1the
=.
(PPPii (PPPPifi)1)=P==i P1i)1Q)=Q
. i 2.xii 11 ( k+.A )+x ii f ( k.A)
Q
k.)A+)i k1i .A
1 ( k.A )i i
( i i 1 )
2 .2 (2k .(Ak .)(A
2 ( k.A )ii 11 i (( k.A ))ii
x
Q x
(Pi Pi 1 ) = . i 1 + i
2 ( k.A )
( k.A)
Q x i1 x i

i 1
(Pi Pi1Q1 )= =1 1. 2 . 2 2 + . iP P
. 22xi (1k.A )i.1
(P.PAi(i1))i Pi 1i)1 ) Darcy Law:
xi i(k.single-phase
Q = Q1=to.the
x i 1 form
ofx.((iP
which rearranges
the
+
Q = . xfollowing

P
x

+x i ( k.Ai ) Pi 1 )
.+)A )
k+
x ii(11k(.A
i

1 i ( k.A )i i
1 ( k.A )i 1 2 i(k1 .A
)
Q = . ( k.A )i 1 ( k.A )ii .( Pi Pi 1 )
x i 1 + x i

1
2
Q = . ( k.A )i 1 ( k.A )i .( Pi Pi 1 )
x i 1 + x i
( k.A )i 1 ( k.A )i
1 i 1i 1Pf
f Pi 1i 1
k i 1i.A
In block (i-1): QkQ

1 .A
i 1 Pf . P.i 1
=i =

(2)

(3)
(4)

(5)

(6)

(7)

But, taking kA as the correct average, then, by definition, the single-phase Darcy
Law is of the form:

kA ( P Pi 1 )
Q= . i
x
( xPi 1 +Pi
kA
1 )i
Q= . i
xi 12+ xi
2

(8)

Comparing identical terms in equations 7 and 8 above gives that:

kA
xi kA
1 + xi
xi 12+ xi
2

2
=
xi 1 2+ xi

=
xi)i11 (k. A
xi)i
(k. A
+

( k . A) i 1 ( k . A) i

which easily rearranges to:

xi 1 + xi
kA =

xix1i 1 + Heriot-Watt
x
Institute of Petroleum Engineering,
+ xi i University
kA =
xi)i11 (k. A
xi)i
(k. A
+

(9)

21

kA ( Pi Pi 1 )
.
xi 1 + xi
2 ( T (So ))i 1/ 2 or ( o (So ))i +1/ 2

x i1 + xi

kA = x x

i 1
i
+

kA
2 p P
p Pi Pi 1
(k.A=)
(k.A)

xpi =1 ikA. xBi . x = kA. B . (10)


xi 1 +xi i1 Q
p x i 1 + x i
+ p

2
( k . A) i 1 ( k . A) i

Q=

Note: This is an important result and in particular, we observe the following:

(i) The appropriate


average kA is not the arithmetic average, it is the harmonic
average weighted
bythe
x grid
+ xblocksizes.
i 1
i
kA =
xi 1
xi
average+gives
more weighting to the lower permeability
much
(ii) This harmonic
k. A)i 1 (kp. A)i
(

value. If the grid sizes are equal, this reduced to the exact harmonic average, k ,
H

of the permeabilities as follows:

( (S ))

Bp
1
1 1
1
=
+
2 ki 1 ki
kH
x

i 1 / 2

or ( o (So ))i +1/ 2

(11)

3.9P
Pi Pi 1
This can be seen for the following example: k1 = 200 mD, k2 = 2 mD kH =
krp
Q p = kA. p . = kA. p .

mD. We would
expect the flowsxto be+much
by the
x
xi T (more
p =
Blower
Bp x i 1 + x i
p
So ) i strongly
or affected
i 1
o ( So ) i + 1 / 2

=
1
2
/
permeability since,
p if one of the permeabilities were zero, then the flow would be

2
2
zero, no matter how large the other permeability was.

+B

Pi Pi 1

Bp x i 1 + x i

i 1
i
kBrpp =overi 1into ithe averaging of kA for multi-phase
andcarry
(iii) Theabove
arguments
p =
2
2 Q = kA. p . P = kA. p .
flow.
p
p

Bp x
p

EXERCISE 1.
1. For the two grid blocks below, calculate

kA (in mD.ft.2)

Bp
15 ft.

20 ft.

( (S ))
T

i 1 / 2

or ( o (So ))i +1/ 2

x S
( T ( o ))i 1/ 2 or ( o (So ))i +1/ 2

P
Pi Pi 1
15 ft.
k2
25 ft.
Qi p1 +
=
xkA
. p . = kA. p .
x

B
k1

i
p

or ( o (So ))i +=1/ 2


( T (So ))i 1/ 100
Bp x
Bp x i 1 + x i
2
ft.
2
p P
p Pi P2i 1
120 ft.
Q p = kA. .
= kA. .

x
Bp x

Bp x i 1 + x i

k rp

2
kA
pthe
answer
pcalculated
Pi Pi 1
(i) For k1 = 200 mD and k1 = 185 mD. Compare
P with

Q p = kA. .
= kA
p .
as the arithmetic average.

.
x + x
Bp x i 1 + x i
= i B1 p i x

2 answer with kA
2
(ii) For k1 = 200 mD and k1 = 5 mD. Compare the
pcalculated
as the arithmetic average.

2. If k1 k2, and A1 A2, show that


average.

k rp
kA is approximately
equal to the
pBarithmetic
p
p
p

Bpx

Bp

x x

22

+ x

PP
( T (So ))i 1/ 2

kAkA . ( i i 1 )
2
Q=
=
xxi
xi
xi 1 +xixi 1 +
i 1

S
or
(
)
+
(
)

T
o
2 k. A i 1/ 2 k. A( o(So ))i +1/ 2 Q = kA. p . P = kA.
2
( )i 1 ( )i
p
B x

Gridding And Well Modelling

Qp =

kA. B

p Pi Pi 1

kA
2 p P
= +p =xi kA. . = kA. .

xi 1 Bxi x
xi =1 + xixi 1Q
kA
kA
Bp x i 1 + x i
x
+ p

x
i 1

+(k. A)ii 1 (k. A)i


2
( k . A) i 1 ( k . A) i
kA
3.3 Averaging of the Two-Phase Mobility Term, p
To recap, the single-phase
kA term is taken as a weighted harmonic average and this

1
1
1
1 averages to take for the two-phase flow term, p .
leaves us with the
issue of what
= xi 1++ xi
kA
=
In fact, it is
from the viscosity
p
2
relative permeabilityBterm
ki xi the
kHconvenient
kxii to11 separate
+p as follows:
separately

and consider these


( k . A) i 1 ( k . A) i
Bp

x
k
p = rp
Bp
1 p1 1
1
(12) x
=
+
x
x

+
i
i
1
2 ki 1 ki
kH
=
2 variable than
We do this since the relative
x permeability of phase p is far more
+
+

B
B
i
i
i
i

1
x i 1 + x
the phaseviscosity.
If fact,
that
andwithout
Bp = further discussion, we will simply note =
p =
2
2 are very accurately calculated as the 2
the viscosity and
volume factors
krp formation
k rp
=
x
x

+
arithmeticaverages,
since
they
usually
do
p
i 1
i not vary very much from block to block
=
p
p
i.e. the averages between grid blocks
2 (i-1) and i are:
k rp
p
+ i k
Bi 1 + Bi
rp B =
p = i 1
and
p
2
2
(13)
p

Therefore the situation is summarised in Figure 22:

Qp

Figure 22
Which average should be
taken for the relative
permeability of phase p in
the averaging of block to
block flows?

Qp = - kA.

Harmonic average

krp
p.Bp

Which
average??

Arithmetic averages

We will determine which relative permeability average to take by considering the


physical situation of two-phase oil/water flow from block i (i+1) as shown in
Figure 23. Consider the situation in Figure 23 where:
Block i is at Sw = (1-Sor) i.e. only water can flow; krw > 0 and kro = 0;
Block (i+1) is at Sw = Swc i.e. only oil can flow; krw = 0 and kro > 0.

Institute of Petroleum Engineering, Heriot-Watt University

23

Sor

1 - Sor
Sw

krw > 0

krw = 0

kro = 0

kro > 0

Swc
Block i

Block i + 1

Physically, it is clear in Figure 23 that water can flow from i (i+1) but oil cannot.
Therefore, for flow in this direction, the average water relative permeability must be
non-zero but the average oil relative permeability must be zero. Let us consider the
different averages that are possible in turn:
Harmonic Average: First consider if the harmonic average can be used since this
was appropriate for the single-phase permeability averaging.
Harmonic average of water relative permeabilities = Harm. Av. {krw i >0; krw i+1 = 0}
=0
Likewise, for oil, Harm. Av. {kro i = 0; kro i+1 > 0} = 0
Thus, the harmanic average gives zero flow for both water (incorrect) and oil
(correct). Therefore, it cannot be the harmonic average which is correct for the
relative permeability.
Arithmetic Average: Now consider if the arithmetic average can be used since this
is a natural thing to try and it is certainly the simplest.
Arithmetic average of water relative permeabilities

(k

> 0) + ( k

rw i
rw i +1
= ( krw i > 0) + ( krw i +1 = 0) > 0
2

= 0)

> 0

So, the arithmetic average could be physically correct for the water phase since it
gives the average krw > 0 and thus allows water to flow.

krw > 0

But, for oil the oil phase, we find that : Arith. Av. {kro i = 0; kro i+1 > 0} > 0 and this is
physically incorrect, since oil cannot flow from i (i+1).
Therefore, it cannot be the arithmetic average which is correct for the relative
permeability.
What average does this leave? We simply state the answer and then give some
physical justification.
Upstream Value: In fact, it turns out that the physically correct value of the relative
permeability is simply the upstream value. The upstream value refers to the block from
24

Figure 23
Flow of water from
block i (i+1) and the
corresponding relative
permeability values for
water and oil.

Gridding And Well Modelling

which the flow is coming i.e. in the flow from left to right in Figure 23. This would
be block i. This can be seen to be consistent with the physical situation since:
Flowing from i (i+1), the average water relative permeabilities = Upstream
{krw i} > 0, as required.
Likewise, flowing from i (i+1), the average oil relative permeabilities = Upstream
{kro i } = 0, as required.
Now reversing the flows from (i+1) i, we find that only oil should flow and this
is again consistent since:
Flowing from (i+1) i, the average water relative permeabilities = Upstream
{krw i+1} = 0, as required since water cannot flow in this direction.
Likewise, flowing from (i+1) i, the average oil relative permeabilities = Upstream
{kro i+1} > 0, as required since only oil can flow in this direction.
The situation is summarised in Figure 24.

Qp

Figure 24
The correct inter-block
averages for all terms in
the two-phase block-toblock flows in a reservoir
simulator.

Qp = - kA.

krp
p.Bp

Harmonic average

P
x

Upstream
value of krp

Arithmetic averages

WELLS IN RESERVOIR SIMULATION

4.1 Basic Idea of a Well Model

The only way fluids can be produced from or injected into a reservoir is through the
wells and we must therefore include them in our reservoir simulation model. As you
may know, the area of Well Technology is vast and in addition to the long wellbore
between the reservoir and the surface, there are many other technical features of wells
that can have a major impact on the flows into and out of the reservoir. For example,
there will be safety valves at the surface and many different types of completion in
the well construction itself. Here, we will simplify things as much as possible in
order to extract the central functions of the well that we will have to model in the
simulator. A schematic of the total well is shown in Figure 25 where the details of
the near well formation are shown inset. The near wellbore flows are thought to be
radial in an ideal vertical well and this will have some relevance in modelling the
near-well pressure behaviour, as discussed in Chapter 2 and elaborated upon below.
In addition to these near-well pressure drops, there are several other identifiable

Institute of Petroleum Engineering, Heriot-Watt University

25

pressure drops between the fluids in the reservoir and the surface oil storage facilities
and we may have to model at least some of these. Indeed, it is this topside pressure
behaviour that links or couples the surface with the pressure and flows that we
are trying to model in the reservoir using reservoir simulation. The main decision is
to determine how much of the formation to surface well assembly we will actually
have to model. The main pressure drops are shown in Figure 26 (based on Figure
25 of whole well + Ps) and are associated with:
(i) Formation wellbore flow, Pfw: where fluids flow from a drainage radius,
re, at pressure, Pe, to the wellbore. Figure 26 shows the near-well pressure profile,
in the near-sandface region with bottom hole flowing well pressure (BHFP), Pwf.
Thus the formation wellbore pressure drop, Pfw , is:

Pf w = Pe Pwf

(14)

(ii) The pressure drop, Pwell, that may occur along the completed region of the
= Pthe
Pr s of the well (or the toe of a horizontal well) to the
atm +
wellborePwf
from
bottom
wellbore just at the top of the completed interval. In very long wells, this pressure
drop along the wellbore due to friction may be quite significant although there will
Pe bePwfignored;
o = PI
be casesQwhere
it. can

Well head
Q

o max .

Surface facilities
(separator etc...)

= PI .Pe

To storage /
export

P(r ) =

r
Q
ln
2 (k.h) rw

Well tubulars

P(re ) = Pe Pwf

r
Q
=
ln e
2 (k.h) rw

2 (k.h)
. Pe Pwf
re
. ln
Reservoir showing
rw

Q=

two geological layers

PI =

26

2 (k.h)
r
.ln e
rw

Well

Well completed
in reservoir

Fluid flow from


reservoir layers
to wellbore

Figure 25
Schematic of the fluid flows
into a well in a grid
block model of the reservoir
through to their storage or
export from the field.

Gridding And Well Modelling

Pwh -> export

Surface facilities
(separator etc...)

Well head

To storage / export

Well tubulars

Near wellbore formation


to wellbore Pf -> w

Well

Pe

Pr -> s
pressure drop
from reservoir
top to surface

Figure 26
Schematic of the fluid flows
from the well through to
storage or export showing
the associated pressure
drops that occur in the
system.

Pf -> w
Pwf
rw

re

Fluid flow from


reservoir layers
to wellbore

Reservoir showing
two geological layers

Well completed
in reservoir
Pwell
pressure drop along the well
within reservoir section

(iii) Reservoir surface pressure drop, Prs : the pressure drop from the well at
the top of the completed formation just above the reservoir to the wellhead which is
at pressure, Pwh . This P is quite significant and, locally in any sector of the well,
there will be a local pressure drop vs. flow rate/fluid composition relationship. This
may be calculated from models (often correlations) of multi-phase flow in pipes.
As the fluids move up the wellbore, the pressure drops in oil/water production and
free gas may also appear; thus, we can have three phase flow in the well tubular to
the surface and we may have to incorporate this flow rate/pressure drop behaviour
in our modelling.
(iv) There will frequently be further pressure drops as the fluids flow from the
wellhead through the surface facilities such as the separators, various chokes, etc.
We will not consider this in detail here although it can be an important consideration
in some field cases e.g. if the well is feeding into a network gathering system which
other wells are also feeding into. This could be a complex surface gathering system
network or a multiple-well manifold of a subsea production system.
In this section, we will mainly focus on the formation to wellbore pressure drops. Thus,
our main task is to either calculate or set the well flowing pressure (Pwf) although we
will return briefly to the issue of calculating the pressure drops between the reservoir
and the surface in the discussion below.
To set the scene in modelling wells in a simulator, we will first consider a very
simple model well producing only oil into the wellbore. How do we decide what
Institute of Petroleum Engineering, Heriot-Watt University

27

the volumetric flow rate, Qo, of this well is? Indeed, do we decide or is it set for us
by the reservoir and well properties? We will start with the simplest case where we
basically take what we can get by drawing the wellhead pressure, Pwh, down as
low as possible. Suppose that we simple open it up such that the oil pressure drops
essentially to atmospheric. There is then the additional reservoir to surface pressure
drop, Prs, to consider. Thus, the well flowing pressure, Pwf, would be given by:

Pf w = Pe Pwf

Pwf = Patm + Pr s

(15)

So, what happens ? Clearly, if this value of Pwf > Pe (the reservoir pressure), then no
Pwf
oil can beQproduced.
if Pwf < Pe, then some oil will flow into the well and
o = PI . Pe However,
we can now calculate how much. As we will see, this will depend on the physical
properties of the system such as the permeability of the rock, the viscosity of the oil,
.PethePwell etc. However, in our simple conceptual well, we will
o max . == PI
the preciseQ
P
Pof
geometry
e
f w
wf
take all of these quantities as givens for the moment. Suppose the well does flow
at a volumetric flow rate, Qo, for reservoir pressure, Pe, and well flowing pressure,
r
Pwf . We can
index, PI, of the well as follows:
=r )P=atmdefine
+QPar productivity
PwfP(then
sln


2 (k.h) rw

(16)
(
)
r
Q
P(r ) = ( P P ) =
ln
where possible units of PI could2
be( kbbl/day/psi,
.h) r for example. The above equation

Qo = PI . Pe Pwf
e

wf

= PImuch
.Pe oil is produced per psi of drawdown. This simple equation
Qo max .how
basically states
takes us back to our original question on what/who decides on Qo?. In our simple
case, the
answer
.he ) clear;
Pf w2=is(know
P
Pwf i.e. some things are givens - e.g. PI and Pe in a
=
.P
e-
Prwfsome

virgin oil Q
producing
system
and
we can set within limits - e.g. Pwf by setting
P(r ) = re ln
. ln2 (kHowever,
the wellhead pressure.
you
may
be able to set a flowrate, Qo, by installing
.h) rw
r

w
a downhole
ESP
= Patm(an
+
Pr -s electrical submersible pump). In that case, you would
Pwfpump
set Qo and then calculate Pwf where we are still considering the reservoir pressure
(Pe) as a given. But clearly we cannot
any arbitrarily large value since the
Q set Qo rto
e
2

.
k
h
(
)

=
ln
P
r
P
P
(
)
lowest possible
value
vacuum),which
would then set a maximum value
PIo = ePI
Q
PPwfwf = 0 (a
. Peeof
2 (k.h) rw
re wf
of Qo given by:

( )

((

))

.ln
rw
(17)
Qo max .2= (PI
k.h.P)e
Q=
. Pe Pwf
re set either a pressure or a flowrate but (a) not both and (b)
So, in summary,
. we
ln can

in either case, within


limits.
rQ
r

w
P(r ) =
ln
2 (k.h) rw

But, cant we affect the well PI or the reservoir pressure, Pe? We can actually affect
(stimulating
k.h)
the PI of a well 2by
it possibly by locally hydraulically fracturing the
PI =
well or by acidising it rtoincrease the effective permeability of the near well region.
e
re
Q
.lnPincrease
In addition,
more commonly
Pwf (or
Pwe
=

=
ln
(re )can
maintain) the reservoir pressure
e
rw
2 (k.h) to some
rw extent by injecting a fluid - usually
(which relates to the reservoir
energy)
water or gas in another injector well. However, the basic well controls are either
setting pressure or flow rate and this must be kept in mind when we model wells in
2 (k.hWe
) elaborate on these ideas in the following section where we
reservoir Q
simulation.
=
. Pe Pwf
introduce the central
ridea
of a well model.
e

. ln
rw

28

PI =

2 (k.h)

(
)
= (P P )

Gridding And Well Modelling

Pf w = Pe Pwf
Pf w

wf

Pwf = Patm + Pr s
= Patm +for
PSingle
PwfModels
rs
4.2 Well
and Two-Phase Flow

(
(

)
)

We now consider how a well model can be developed, firstly in our simple conceptual
Qo = PI . Pe Pwf
reservoir producing
only oil. Figure 27 shows the local pressure profiles in a simple
Qo = well
PI . Psystem
homogeneous
e Pwf in single phase flow (see Chapter 2, section 3.5). The
pressure profile close to the wellbore, assuming radial flow, was derived in Chapter
Qo max . = PI .Pe
2 and is given
by (section 3.5):

Qo max . = PI .Pe

r
Q
ln
2Q(k.h) rrw

=
P
r
(
)
Pf w = 2Pe(k.hP)wfln r
P(r ) =

(18)

(
)
Taking theP
pressure
P radius
= (at
P )r as being the reservoir pressure, P , then gives:
r
Q

=
ln
P
r
P
P
(
)
(
)

P = P + P
2Q(k.h) r
P(r ) = ( P P ) =
ln
P = P + P
2 (k.h) r
(19)
Q = PI
2.((kP.h) P )
Q=
.( P P )

r
Q
=
PI
P

.
P
k
h
2

.
(
)
Q
(
(a)
(b)
Q = . ln .( P) P )
Grid block
r

Well at BHFP
P
w

f w

wf
wf

Figure 27
Schematic showing (a)
the near well pressure
profiles that occur in a
radial system and (b)
corresponding quantities in
a Cartesian grid block

e
e

atm
atm

wf

wf
rs

wf
rs

wf

ee

wf

Qo max. .=lnPI .wePe



r.wP
= PI
Q
o max .

w
e
w

wf

wf

pressure

Pwf

2 (kQ
.h)
r
PIP(=r )P(r)
= P(r) - ln
Pwf

=
h
22(
k.r(hek).h) r
rw
PI = .lnQ
rw ln
Pwf P(r ) =
.ln2 (ek.h) rw
x
rw
r rw
re
r
Q
P(re ) = Pe Pwf =
ln e
2Q(k.h) rrwe

=
ln
P(re ) =be P
P
which caneasily
arranged
to
obtain:
e
wf
2 (k.h) rw

(
(

)
)

Pwf

y
(assume x = y)

2 (k.h)
. Pe Pwf
k.rhe )
2

(
Q = . ln . Pe Pwf
(20)
rrw
. ln e
rw
and hence from equation
16 above, we can identify the productivity index
2 (k.h)
(PI) of the well as:
PI =
2 (k.rhe )
PI = .ln
rrw
.ln e
(21)
rw
Q=

(
(

)
)

This now demonstrates exactly how the quantities k, h, , rw and re affect the well
productivity. All of these factors behave as we might expect them to physically e.g.
as k, PI; as , PI etc.
Now consider how this relates to the pressures in the simulation block shown in Figures
27 and 28. In a grid block, the pressure is thought of as being constant throughout
the block although we know that it should be varying continuously across the block;
we will refer to this as the average block pressure, P . The size of the grid block in
Institute of Petroleum Engineering, Heriot-Watt University

29

our example is (x, y) and, for simplicity, we will assume that, x = y. Looking
at the expression for PI in equation 21 and the quantities we have in the grid block,
it is easy to make direct relations for some of them - obviously k, and h and also
possibly rw and Pwf , although these latter two do not seem to appear in the grid model.
The drainage radius, re, and the reservoir pressure, Pe, which appear in the radial model
do not appear in the grid model - instead, the block size (x, y) and average block
pressure ( P ) appear. This immediately suggests the following 2 questions:
1. what is the relation between re , and the block size (x, y)?
2. what is the relation between Pe and the average grid block pressure, P ?
Well at BHFP
= Pwf

Grid block
pressure
P

re

Pwf

re

=>

e =

h
y

Pe

How do we choose
re such that

Pe =P ??

Pwf

Relation
re <-> x, y ??

rw

re

(assume x = y)

re
rw

re is where P(re) =P
and re = re (x, y) but
what is the formula ?

Figure 29
The relation between re and
the block dimensions, x
and y.

The issue is defined quite clearly in Figure 28. From this figure, we would like to
choose re such that Pe coincides with the average grid block pressure, P . The latter
quantity ( P ) is calculated in the simulation itself. In fact, we need to know how to
calculate re from the quantities x and y as indicated in Figure 29 where we show
the re as function of x and y, i.e. re(x, y).
If we know the formula for re(x, y), then we can calculate the PI (equation 21) and
use this inPthe simulation to couple the quantities Qo (oil flow rate) and P (average
grid block pressure); i.e.

Q o = PI.( P Pwf )

30

r 2( k.h ) ( P( r ) Pwf )
ln =
.

Q
rw

Figure 28
Schematic indicating how
the near well pressures
relate to the corresponding
quantities in a Cartesian
grid block

(22)

Gridding And Well Modelling

Here, we can set either Qo or Pwf and then calculate the other one from P (and the
known PIP). This was achieved in a very simple but ingenious way in a classic paper
by Donald Peaceman (1978), another pioneer of numerical reservoir simulation.
Peaceman did this by carrying out a 2D numerical solution of the pressure equation
on a Cartesian
. Pgrid
Pfor
Q o =(x,
PIy)
wf a quarter five-spot configuration as shown in Figure 30.
But, from equation 19, we know that:

r 2 ( k.h ) ( P( r ) Pwf )
ln =
.
rw PI
Q
.( P Pwf )
Q o =

(23)

Hence, if we plot the pressure at grid blocks away from the well block vs. the well
re r 0.2 x
( r ) asPshown
on a logarithmic
block spacing
in Figure 31, then we can extrapolate
2 ( k.h ) Pscale
wf
ln
=
.

back to find the


equivalent
radius
where
=
P
in terms of the well block dimension
rw

Q P
e
(x). It turns out that the2 simple2 relation is (for x = y):

P
re 0.14 x + y

re 0.2 x
Q oo = PI.o(.P( PPPwfwf) );

Q w = PI w .( P Pwf )

(24)

Therefore, we have a very


simple
way of calculating the PI or well connection
2
2
r
x
0
14

.
y
e
factor as it is sometimes called of a well in a simulation grid block.

r 2 kh
2 .k( kro.(hS)o )( P( r ) Pwf )

ln
PIPeaceman
o = = formula. applies to a well in a radial environment (the five-spot
The simple
rw PI
. P Pre ; QQ = PI . P P
Q o =
w to radial
w
wf a common 2D Cartesian grid)
configuration
is asoclose
aswfwe
using
can get
o .ln
r

and for x = y. In fact,wsome modification to the simple formula is required for


wells in corner
or set close to a boundary and these are shown in Figure
re 0.locations
2 x
2
khbut
.k rothe
So well
(
) is isolated (radial flow), then:
32. Also,PI
if x

y,
=

o
2kh.k (S )
PI w = .ln rwre w
o
re 0.14
x2rwr+ey 2
w .ln

rw

Figure 30
The 2D areal grid used to
compare pressures with
the expected radial profiles
(from Peaceman, 1978)

(25)

Q o = PI2
.kh
(P.k P(S); ) Q w = PI w .(P Pwf )
PI w = o 11 rwwf w
10 re
w .ln
9

w )
2 kh.8k ro (rS
o
7
PI o =
6 r
o .ln5 e
4 rw

PI w =

3
2
2kh.10k rw Sw
-1 r

( )

w .ln-1e 0 1
rw

2 3 4 5 6 7 8 9 10 11

Institute of Petroleum Engineering, Heriot-Watt University

31

0.6
On Edge: i = 0 or j = 0
On Diagonal: i = j
ij0

0.5

pij - po
q / kh

0.4

(2.2)
(2.1)
(2.0)

0.3

(1.1)
(1.0)

0.2
0.1

r Ao / x From areal average pressure

ro / x

0
0.1

0.2

0.4

0.6

5 6

Figure 31
Log plot of the pressures
from the 2D areal grid used
to find re (from Peaceman,
1978)

Factor in terms of (re / x)


(re / x) = 0.2
(Peaceman's equation)

(re / x) = 0.196
implies no flow boundary

(re / x) = 0.433

(re / x) = 0.193

(re / x) = 0.72

For anisotropic
permeabilities
kx ky

32

re = 0.28

ky
kx

1/2

.x +

kx
ky

1/2

.y

ky 1/4
k 1/4
+ x
kx
ky

1/2

Figure 32
Well factors for wells in
various positions relative
to the boundaries; after
Kuniansky and Hillstad
(1980)

Gridding And Well Modelling

We may find that, in a given simulation of a field case, that we input all the known
or estimated data but the well in our simulation does not perform like the real case.
It may produce more (higher PI) or it may produce less (lower PI) than expected.
The former case may be due to the well being stimulated and the second case may
be due to well damage. Within limits, we may adjust the calculated well PI in the
simulation model in order to reproduce the observed field behaviour. However,
we should think carefully before making such changes since the simulated well
productivity may be wrong because some (or several) other aspects of the reservoir
simulator input data are wrong.
It is now relatively straightforward to extend our discussion on PI and simple well
models in a homogeneous single layer system to the flow of two phases - say, oil
and water - as shown in Figure 33. Since two phases are being produced, then the
saturations of both oil and water (So and Sw) must be at values where their relative
permeabilities are > 0 (i.e. = > So > Sor ; Sw > Swc).
Well radius = rw

Qw

Qo

P
P
Q o = Saturation
PI.( P Pwf )

Figure 33
Near well two-phase flows
in a Cartesian grid block.

w .and
Q o = SPI
P SoPwf )
(
Q o = PI.( P Pwf
h )
r 2 ( k.h ) ( P( r ) Pwf )
ln =
.
rw 2 (k.h ) (xP( r )Q Pwf )
ln r = 2 ( k.h ) . ( P( r ) Pwf )
.
Q
ln rw =

Q
rw
(assume
x = y)
re 0.2 x

r 0.2 x

e
We now apply
were developed above for the single phase case.
re the
0.2 same
x 2ideas as
2
From the rradial
two phase Darcy Law, the volumetric production rates of oil and
e 0.14 x + y
water are given by:
2
2

re 0.14 x 2 + y 2
re 0.14 x + y
Q o = PI o .( P Pwf ); Q w = PI w .( P Pwf )

(26)

Q o = PI o .( P Pwf ); Q w = PI w .( P Pwf )
Pthe
PI oPI
.(wPare
Q w water
= PI wproductivity
.( P Pwf ) indices, respectively, and
where theQPIo 0=and
wf );oil and
2 kh
.k ro (So )
PI o =
2 kh
.kror(eSo )
PI o = 2
o .ln
kh
.kror(So )
rw
PI o =
o .ln re
o .ln rwe
(27)
r
2kh.k rww(Sw )
and
PI w =
2
.krwr(eSw )
.ln
kh
PI w = 2
w
kh.krwr(Sw )
rw
PI w =
w .ln re
w .ln rwe
rw
(28)
Institute of Petroleum Engineering, Heriot-Watt University

33

where, again, re is calculated using Peacemans formula. In the above equation, we


have not incorporated the separate phase pressure, Po and Pw, in the well block but
these may be used in a given calculation.

4.3 Well Modelling in a Multi-Layer System

The most common case which is modelled is where we have multi-phase (e.g. two
phase oil/water) flow in a layered system where the layers are of different permeability.
This situation is shown for a simple four layer system in Figure 34.
Clearly, there are additional issues in this system since all four layers may be
producing both oil and water and the proportions of each phase may be changing
as the saturations (and hence relative permeabilities) change. In addition, there is
also a gravitational potential in each layer which we may have to take into account.
Using the notation in Figure 34, we note that the oil flows in layer k (k = 1, 2, 3, 4,
in this example) are:

Q ok = PI ok ( Pk Pwfk ) =

2 ( khk ro (So ))k

(Pk Pwfk )
rek
o .ln
2 ( khk
))
ro (rS
w o k
(29)
Q ok = PI ok ( Pk Pwfk ) =
(Pk Pwfk )
rek
o .ln
and a similar expression
applies for
water.
3
3
rw The total oil and water flows in
the well are:
Q T = (Q ok + Q wk ) = ( PI ok + PI wk )( Pk Pwfk )
k =1
3

k =1
3

Q T = (Q ok + Q wk ) = ( PI ok + PI wk )( Pk Pwfk )
1
k =1
Q Tk =k(=Q
ok + Q wk ) = ( PI ok + PI wk )( Pk Pwf )

(
( )
)

(30)

where we have taken the mean block pressure and also the well flowing pressure in
Q Tk = (Q ok + Q wk ) = ( PI ok + PI wk ) Pk Pwf
layer k asQbeing
the same for both oil
water phases. Again, in the layered case,
Pwf
and
i ,khk
j, k ro (So )
o i , j, k = PI o i , j, k . P i , j, k 2
k
we can specify
the flowing bottom
hole
and then we can
Q ok = the
PI oktotal
Pk flow
Pwfkor =
Pk pressure
Pwfk

r
calculate the other one using the above
(In fact, we can also specify
ek
.lnrelations.
find the flow of the other and the
P ior
Q oeither
PIQ
, j, k Pwf i ,oj, k
o i , j,-k .oil
i , j, k =B
the flow of
phase
water
and
then
r

o
o
w
Q wflowing
+ Q w For example, suppose
inj =
bottom hole
we specify the total flow, QT.
Bpressure).
w
We need to decide how this total flow is made up - i.e. what are the separate Qo and Qw
3 Bo Q o
(QT = Qo +QQ
) and
how this
is3allocated from each of the layers in the system.
=
+QQflow
w
w
inj
w =
=
+
Pwfkwell flowing pressure,
Q
Q
PI wkis) P
(
)
(PIthat
T
ok the assumption
wk
ok +there
BQ
For simplicity,
weBmake
a ksingle
w
o
k =1 o
k =1
Q
Q
>
+
Pwf (i.e. Pwf1w=injPwf2 =BPwf3 = Pwf4w). Hence, for each layer, k :

(
(

B Q
Q wTkinj=>(Q oko +oQ+wkQ) w= ( PI ok + PI wk )( Pk Pwf )
Bw

(31)

Everything is known in the above equation which allows us to determine the allocation
Pwfcan
Q o from
PI o i , layer
, j, k we
j, k . P i and
i , j, k calculate the corresponding bottom hole
i , j, k = each
of all fluids
flowing pressure, Pwf.

Bo Q o we could specify the well flowing pressure, P , and then


In a very Q
similar= manner,
+ Qw
wf
w inj
Bw flows, Qok and Qwk etc. in each layer.
calculate the individual

34

Q w inj >

Bo Q o
+ Qw
Bw

Gridding And Well Modelling

Well
completion

(a)

Layer
1
2
3
4

(b)

QT = Qo + Qw
Qo1
Qw1 k1, h1, P1, Sw1
Qo2
k ,h ,P ,S
Qw2 2 2 2 w2

Figure 34
Well modelling of twophase flow in a multilayered system

Qo3
Qw3 k3, h3, P3, Sw3

Q ok = PI ok ( Pk Pwfk ) =

2 ( khk ro (So ))k

r
o .ln ek
4.4 Modelling Horizontal Wells rw

(P

Qo4
Qw4 k4, h4, P4, Sw4

Pwfk )

Figure 35 shows the trajectory of a horizontal well in a reservoir simulation


model. This is not well represented by the purely radial r/z model grid discussed
3
3
in Section 2.1 above in the context of a vertical well. Hence, it is less likely that
Q T = (Q ok + Q wk ) = ( PI ok + PI wk ) Pk Pwfk
the well connection
factors calculated
as shown in previous sections will apply for
k =1
k =1
a horizontal well. This is broadly true although the basic principle is very similar.
That is, each well sector intersects a grid block (i,j,k) even although the well may
be Q
going
through this block in say the x(i) direction and the flows between the well
Tk = (Q ok + Q wk ) = ( PI ok + PI wk ) Pk Pwf
sector and the grid are given by an expression of the form:

Q o i , j, k = PI o i , j, k . P i , j, k Pwf i , j, k

(32)

where the actual


of the productivity index expression, PIoi,j,k , may be rather more
Bo Qform
o
Q w inj =
+ Qwell
complex
since (a) the
may intersect the block in a more complex way and (b)
w
B
w
the aspect ratio of the block is rather different when a horizontal well intersects it in
that the x-direction well is very close to the z-boundaries since z is often smaller
than x or y.
B Q

Q w inj >

Bw

+ Qw

Institute of Petroleum Engineering, Heriot-Watt University

35

Figure 35
Cartesian grid cut from a
3D reservoir model showing
two horizontal wells going
through the system; two
vertical wells also shown.

4.5 Hierarchies of Wells and Well Controls

Simple well control can be understood in terms of the well models discussed above.
For a single well, we can essentially set the well flowing pressure and then calculate the
flows or vice versa but not both and with certain constraints (see above). Alternatively,
we may set the wellhead pressure and then calculate the Pwf from the - calculated
or input - well formation to surface pressure drops etc. We now consider controls
on pairs, then groups and then clusters of groups of wells in a field - indeed, we can
even couple together the wells from several reservoirs and set more global controls
and this will be described briefly.
For a simple injector/producer well pair, for example in a 2D x/z cross-section, we
can apply a range of well controls. Suppose this is a simple waterflood in an oil/water
system. One of the most common controls is to fix the water injection rate at the injector
but with (upper) limits on the well flowing pressure. The corresponding producer
is then controlled by setting the bottom hole flowing pressure and then allowing the
calculation of the oil and water phase flows (Qo and Qw). The volumetric production
will be approximately balanced with the total production volume being of the order
of the injected water volume - but not quite the same. Do you know why this is?
Clearly the formation volume factors (Bo and Bw) will affect the exact production
volumes; when we are injecting water and producing 100% oil, the reservoir volume
of injected water per day is Qw.Bw and this will displace (virtually) the same reservoir
volume of oil. The volume of oil produced per day is Qo stb which is actually Qo.Bo
reservoir bbls, equating these reservoir volumes shows that if we inject water at a
rate of Qw (stb/day), we produce oil at a rate of Qw.Bw/Bo stb/day. Since Bo > Bw,
then the volumetric production rate of oil (in stb/day) is lower than the injected water
injection rate (in stb/day). This must be taken into account in considering well control
by voidage replacement as discussed below.
If we wish to set injected Qw to precisely voidage replace whatever is produced, then
we can do so to a good approximation by noting that if the production rate of oil

36

ok

ok

wfk

r
o .ln ek
rw

wfk

Gridding
And Well Modelling
3

Q T = (Q ok + Q wk ) = ( PI ok + PI wk )( Pk Pwfk )
k =1

k =1

Q Tk = (Q ok + Q wk ) = ( PI ok + PI wk )( Pk Pwf )

and water in our simulation is currently, Qo and Qw. What volume of injected water
2 khk ro (So ) k
must Q
we inject
to exactly replace the reservoir
volume
these two phases? This is
Pk Pof
ok = PI ok Pk Pwfk =
wfk

r
now quiteQstraightforward
since
and,
from
the
above
discussion,
it is evident that the
=
.
P

P
PI
,
,
i
j
k
ek
o i , j, k
wf i , j, k
o i , j, k
o .ln
Q (in stb/day) is given by:
quantity of water that must be injected,
rw inj

B Q
Q w3inj = o o + Q w 3
B
Q T = (Q ok + wQ wk ) = ( PI ok + PI wk )( Pk Pwfk )
k =1

(33)

k =1

Hence, we would gradually adjust the volume of water injected in the simulation
Bo Q o
Q w inj
+ Q wjust produced (at the last time step say) to the above
model based
on>what
we have
Bwwk ) = replace.
Q Tkin=order
Q
Pk Pwf the most common option would
(Q okto+voidage
(PI ok + PI
quantity
Atwk
the) producer,
be to constrain by bottom hole flowing pressure as described above.

P i ,less
Pwf i , j, k to constrain all wells by volumetric injection/
PI o i , j, k .but
Note Q
that
j, k common
o i , it
j, kis=possible
production rate. We can see why if we consider an incompressible fluid where it
is clearly impossible for the injection and production rates to be different since the
B go
Q to + or - , depending on whether we over- or under-injected,
pressure would
Q w inj = o o + Q w
respectively.
Although
it is possible to specify different volumetric flow rates at
Bw
injector and producer for a compressible fluid, this can only be done within very tight
limits and the pressures tend to go to unrealistic limits e.g. if we over-inject
Bo Q o

+ Q w ), the pressure tends to rise to unphysically high levels


(i.e. Q w inj > B
w
well above fracture pressure of the reservoir rock.
4.5 CLOSING REMARKS - GRIDDING AND WELL MODELLING
In this section, the student has been presented with a largely non-mathematical
description of gridding and well modelling in reservoir simulation. A more
mathematical treatment of these issues will be given as we develop the flow equations
and consider their numerical solution in Chapter 5 and 6, respectively.

Institute of Petroleum Engineering, Heriot-Watt University

37

Numerical Methods in Reservoir Simulation

CONTENTS
1.

INTRODUCTION

2.

REVIEW OF FINITE DIFFERENCES

3.

APPLICATION OF FINITE DIFFERENCES TO


PARTIAL DIFFERENTIAL EQUATIONS
(PDEs)
3.1. Explicit Finite Difference Approximation
of the Linear Pressure Equation
3.2. Implicit Finite Difference Approximation
of the Linear Pressure Equation
3.3. Implicit Finite Difference Approximation
of the 2D Pressure Equation
3.3.1 Discretisation of the 2D Pressure Equation
3.3.2 Numbering Schemes in Solving the 2D
Pressure Equation
3.4. Implicit Finite Difference Approximation
of Non-linear Pressure Equations

4.

APPLICATION OF FINITE DIFFERENCES TO


TWO-PHASE FLOW
4.1 Discretisation of the Two-Phase Pressure
and Saturation Equations
4.2 IMPES Strategy for Solving the Two-Phase
Pressure and Saturation Equations

5.

THE NUMERICAL SOLUTION OF LINEAR


EQUATIONS
5.1. Introduction to Linear Equations
5.2. General Methods for Solving Linear
Equations
5.3. Direct Methods for Solving Linear
Equations
5.4. Iterative Methods for Solving Linear
Equations
5.5. A Comparison of Iterative and Direct
Methods for Solving Linear Equations

6.

DIRECT SOLUTION OF THE NON-LINEAR


EQUATIONS OF MULTI-PHASE FLOW
6.1. Introduction to Sets of Non-linear
Equations
6.2. Newtons Method for Solving Sets of Nonlinear Equations
6.3. Newtons Method Applied to the Non-linear
Equations of Two-Phase Flow

7.

NUMERICAL DISPERSION - A
MATHEMATICAL APPROACH
7.1. Introduction to the Problem
7.2. Mathematical Derivation of Numerical
Dispersion

8.

CLOSING REMARKS

APPENDIX A: Some Useful Matrix Theorems.

LEARNING OBJECTIVES:
Having worked through this chapter the student should be able to:
write down from memory simple finite difference expressions for derivatives,
(P/x), (P/t) and (2P/x2) explaining your spatial (space) and temporal (time)
n
n +1
n +1
n +1
notation ( Pi , Pi , Pi +1 , Pi 1 etc. ); for (P/x), the student should know the
meaning of the forward difference, the backward difference and the central difference
and the order of the error associated with each, O(x) or O(x2).
apply finite difference approximations to a simple partial differential equation
(PDE) such as the diffusion equation and explain what is meant by an explicit and
an implicit numerical scheme.
write a simple spreadsheet to solve the explicit numerical scheme for a simple
PDE for given boundary and initial conditions and be able to describe the effect of
time step size, t.
show how the implicit finite difference scheme applied to a simple linear PDE
leads to a set of linear equations which are tridiagonal in 1D and pentadiagonal in
2D.
derive the structure of the pentadiagonal A-matrix in 2D for a given numbering
scheme going from (i, j) notation to m-notation where m is an ordered numbering
scheme e.g. for the natural numbering scheme, m = (j - 1).NX + i
describe a solution strategy for the non-linear single phase 2D pressure equation
where the fluid and rock compressibility (and density, , and viscosity, ) are functions
of the dependent variable, pressure, P(x,y,t).
write down the discretised form of both the pressure and saturation equation for
two-phase flow given the governing equations (in simplified form in 1D), and be able
to explain why these lead to sets of non-linear algebraic equations.
outline with an explanation and a simple flow chart the main idea behind the IMPES
solution strategy for the discretised two-phase flow equations.
write down the expanded expressions for a set of linear equations which, in compact
form are written A.x = b, where the matrix A is an nxn matrix of (known) coefficients
(aij; i = rows, j = columns), b is a vector of n (known) values and x is the vector of n
unknowns which we are solving for.
explain clearly the main differences between a direct and an iterative solution
method for the set of linear equations, A.x = b.
write down the algorithm for a very simple iterative scheme for solving A.x = b,
and be able to describe the significance of the initial guess, x(0) , what is meant by
iteration (and iteration counter, ), the idea of convergence of x() as ; and be
able to comment on the number of iterations required for convergence, Niter.

Numerical Methods in Reservoir Simulation

explain how to apply (without derivation) the Newton-Raphson method for


solving a single non-linear algebraic equation, f ( x ) = 0 ; Newton-Raphson scheme
=> x ( +1) = x ( )

f ( x ( ) )
.
f ' ( x ( ) )

extend the application of the Newton-Raphson to sets of non-linear algebraic


equations such as those arising from the discretisation of the two-phase pressure and

S
P

saturation equations; F( X ) = 0, where X = and S and P are the (unknown)


vectors of saturation and pressure; (given the Newton-Raphson expression,

X ( +1) = X ( ) J ( X ( ) ) .F( X ( ) )).


1

understand, but not be able to reproduce the detailed derivation of, the more
mathematical explanation of numerical dispersion.

NUMERICAL METHODS IN RESERVOIR SIMULATION


As noted in Chapter 5, the multi-phase flow equations for real systems are so complex
that it is not remotely possible to solve them analytically. In practice these equations
can only be solved numerically. The most commonly applied numerical methods
are based on finite difference approximations of the flow equations and this approach
will be followed in this chapter.
After a brief review of finite differences, we go on to apply them to very simple
systems such as for the simplified 1D pressure equation (derived in Chapter 5). This
equation does not need to be solved numerically but it demonstrates how explicit
and implicit finite difference solution can be developed. We then show how sets of
linear equations arise in solving implicit equations and we consider solution of the
linear equations in an elementary manner. We then consider how the 2D pressure
equation is solved.
In the numerical solution of the multi-phase flow equations, we need to solve for
pressure and flow. An outline of how this is done is presented but we do not go into
great detail.

1.

INTRODUCTION

At the very start of this course, we considered a very simple simulation model for a
growing colony of bacteria. The number of bacteria, N, grew with a rate proportional
to N itself i.e. (dN/dt) = .N, where is a constant. We saw that this simple equation
.t
had a well-known analytical solution, N( t ) = N o .e , where No is the number of
bacteria at time, t =0. This exponential growth law provides the solution to our
model. However, we also introduced the idea of a numerical model where, even
although we could solve the problem analytically, we formulated it this way, in any
case. The numerical version of the model came up with the algorithm (or recipe)
N n +1 = (1 + .t ).N n . In the exercise in Chapter 1, you should have compared the
results of the analytical and numerical models and found out that, they get closer as
Institute of Petroleum Engineering, Heriot-Watt University

we take successively smaller time steps, t. In this case, we say that the numerical
model converges to the analytical model. In fact, in many areas of science and
engineering, we often apply a numerical technique to a problem we can already
solve analytically i.e. where we know the answer. Why would we do this? The
answer is that we might be testing the numerical method to see how closely it gives
the right answer. More commonly, we might test several - maybe 3 or 4 - numerical
techniques to determine which one works best. The phrase works best in the context
of a numerical method usually means gives the most accurate numerical agreement
with the analytical answer for the least amount of computational work. Note the
importance of this balance between accuracy and work for a numerical method.
There may be no point in having a numerical method that is twice as accurate (in
some sense) for ten times the amount of computational work.
In Chapter 5, we already met the equations for single-phase and two-phase flow of
compressible fluids through porous media. These turned out to be non-linear partial
differential equations (PDEs). Recall that a non-linear PDE is one where certain
coefficients in the equation depend on the answer we are trying to find e.g. Sw(x,t),
P(x,y,z,t) etc. For example, for single phase compressible flow, the equation for
pressure, P(x,t), in 1D is given by:

P k. P
c( P) =

t x x

(1)

In this equation, both the generalised compressibility term, c(P) (of both the rock and
the fluid), and the density, (P), terms depend on the unknown pressure, P, which
we are trying to find. As noted previously, such non-linear PDEs are very difficult
to solve analytically and we must usually resort to numerical methods. The main
topic of this module is on how we solve the reservoir flow equations numerically.
This process involves the following steps:
(i) Firstly, we must take the PDE describing the flow process and chop it up into
grid blocks in space. This is known as spatial discretisation and, in this course,
we will exclusively use finite difference methods for this purpose.
(ii) When we apply finite differences, we usually end up with sets of non-linear
algebraic equations which are still quite difficult to solve. In some cases, we do
solve these non-linear equations. However, we usually linearise these equations
in order to obtain a set of linear equations.
(iii) We then solve the resulting sets of linear equations. Many numerical options
are available for solving sets of linear equations and these will be discussed below.
This is often done iteratively by repeatedly solving them until the numerical solution
converges.
This module will deal successively with each of the parts of the numerical solution
process, (i) - (iii) above.

Numerical Methods in Reservoir Simulation

2. REVIEW OF FINITE DIFFERENCES


Definition:

A finite difference scheme is simply a way of approximating

2
derivatives of a function (e.g. dP , P , P etc.)
2

dx t x

numerically from either point or block values of the function.


The main concept of finite difference approximation is best illustrated by the
following simple example where we refer to Figure 1. Study the Notation in this
figure, since it is the basis of that used throughout this chapter. The main task of the
finite difference approach is to represent the derivatives of the function, P(x), in an
approximate manner
2
dP , P , P etc.
2
i.e. dx t x

First consider how we might approximate (dP/dx) at xi using the quantities in Figure
1. In fact, it is easily seen that there are three ways we may do this as follows:
Approximation 1 -Forward Differences (fd): we may take the slope between Pi and
Pi+1 as being approximately dP at xi to obtain:

dx i

dP = Pi +1 Pi
dx i fd
x

(2)

Approximation 2 -Backward Differences (bd): we may take the slope between Pi-1

dP at x to obtain:
and Pi as being approximately
i
dP = Pi Pi 1
dx i bd
x

dx i

(3)

Approximation 3 -Central Differences (cd): thirdly, we could take the average of


the forward difference (fd) and backward difference (bd) approximations to give us

dP at x . This is known as central differences (cd) and is given by:


i
dx i

dP = 1 dP + dP = 1 Pi +1 Pi + Pi Pi 1
dx i cd 2 dx i fd dx i bd 2 x
x
P P
dP
= i +1 i 1
dx i cd
2.x
(4)

Institute of Petroleum Engineering, Heriot-Watt University

Pi+1
P(x)
Pi
Pi-1
x = constant

x
xi-1

x
xi

xi-1

Notation:
x
= the x- grid spacing or distance between grid points
(i - 1), i, (i + 1) = the x - label (subscript) for the grid point or block numbers
Pi-1, Pi, Pi+1
= the corresponding function values at grid point i-1, i, i+1 etc.

Each of the above approximations to

dP
is shown graphically in Figure 2.
dx i

You may wonder why we bother with three different numerical approximations to

dP
dx i and ask: which is best? There is not an unqualified answer to this question

just yet, but let us take a simple numerical example where we know the right answer
and examine each of the forward, backward and central difference approximations.
The values given in Figure 3 will illustrate how the methods perform. Firstly, simply
calculate the values given by each methods for the data in Figure 3:

dP 3.0042 2.7183 = 2.859 [err +0.14]

dx i fd
0.1
dP 2.7183 2.4596 = 2.587 [err 0.13]

dx i bd
0.1
dP 3.0042 2.7183 = 2.723 [err 0.005]

dx i cd
2 x 0.1

(5)

The quantity in square brackets after each of the finite difference approximations
above is the error i.e. the difference between that method and the right answer which
is 2.7183. As expected from the figure for this case, the forward difference answer
is a little too high (by +0.14) and the backward difference answer is a little too
low (by -0.13). The central difference approximation is rather better than either
of the previous ones. In fact, we note that the fd and bd methods give an error of
order x, the grid spacing (where, as engineers, we are saying 0.14 0.1 !). The
cd approximation, on the other hand, has an error of order x2 (where again we are
6

Figure 1
Notation for the application
of finite difference methods
for approximating
derivatives.

Numerical Methods in Reservoir Simulation

saying 0.005 (0.1x0.1 = 0.01). More formally, we say that the error in the fd and bd
approximations are "of order x" which we denote, O(x), and the cd approximation
has an error "of order x2" which we denote, O(x2).

2 P
2
Now consider the finite difference approximation of the second derivative, x
P

Going back to Figure 1, the definition of second derivative at x xi is the rate of

change of slope (dP/dx) at xi. Therefore, we can evaluate this derivative between xi-1
and xi (i.e. the bd approximation) and then do the same between xi and xi+1 (i.e. the
fd approximation) and simply take the rate of change of these two quantities with
respect to x, as follows:

P
2
x
2

dP dP
dx i fd dx i bd
x
Pi +1 Pi Pi Pi 1
x i fd x i bd

2 P
2
x

( Pi +1 + Pi 1 2 Pi )
x 2

(6)

2 P
2
Calculating the numerical value of x for the example in Figure 3 (where again
2 P
2 = 2.7183 since the example is the exponential function), gives:
x
2 P 3.0042 + 2.4596 2 x 2.7183
= 2.7200 [err 0.0017]
2
x
0.12
(7)
Thus, we can see that the error in this case (err 0.0017 0.12) is actually O(x2).
In the introductory section of Chapter 1, we already applied the idea of finite differences
(although we didnt call it that at the time) to the simple ordinary differential equation

dN
(ODE);
= .N .
dt

Here, N (number of bacteria in the colony) as a function of time, N(t), is the unknown
we want to find. Denoting Nn and Nn+1 as the size of the colony at times t (labelled
n) and t+t (labelled n+1), we applied finite differences to obtain:

Institute of Petroleum Engineering, Heriot-Watt University

n +1
n
dN N N . N n
dt
t

N n +1 (1 + .t ) N n

(8)

This gave us our very simple algorithm to explicitly calculate Nn+1. We then took
this as the current value of N and applied the algorithm repeatedly.
dP
dx i cd

Pi+1 - Pi-1
2.x

Pi+1
P(x)

dP
dx i fd

Pi

Pi+1 - Pi
x

Pi-1

dP
x
xi-1

=
dx i bd

xi

xi+1

dP
= 2.7183
dx x=1.0

d2P
dx2

Pi - Pi-1

x = constant

Figure 2
Graphical illustration of the
finite difference derivatives
calculated by backward
differences (bd), forward
differences (cd) and central
differences.

= 2.7183
x=1.0

Pi+1

3.0042

P(x)
Pi

2.7183

Pi-1

2.4596

0.1
0.9

0.1

1.0

1.1

Figure 3
A numerical example where
the function values and
derivative values are known
(P(x) = ex)

Numerical Methods in Reservoir Simulation

EXERCISE 1.

Apply finite differences to the solution of the equation:

dy = 2. y 2 + 4
dt
where, at t = 0, y(t = 0) = 1. Take time steps of t = 0.001 (arbitrary time units) and
step the solution forward to t = 0.25. Use the notation yn+1 for the (unknown) y at
n+1 time level and yn for the (known) y value at the current, n, time level.
Plot the numerically calculated y as a function of t between t = 0 and t = 0.25 and
plot it against the analytical value (do the integral to find this).
Answer: is given below where the working is shown in spreadsheet CHAP6Ex1.xls.
This gives the finite difference formula, a spreadsheet implementing it and the analytical
solution for comparison.

3. APPLICATION OF FINITE DIFFERENCES TO PARTIAL


DIFFERENTIAL EQUATIONS (PDEs)
3.1 Explicit Finite Difference Approximation of the Linear Pressure
Equation

We have seen in Chapter 5 that the flow equations are actually partial differential
equations (PDEs) since the unknowns, P(x,t) and Sw(x,t) say, depend on both space and
time. As an example of a linear PDE, we will take the simplified pressure equation
(equation 26; Chapter 5) as follows:

k 2 P
P
=
t c x 2

(9)

k
where the constant c is the hydraulic diffusivity, which we denoted previously

by Dh. As we noted in Chapter 5, this PDE is linear which has known analytical
solutions for various boundary conditions. However, we will again neglect these
and apply numerical methods as an example of how to use finite differences to solve
PDEs numerically. To make things even simpler, we will take Dh = 1, giving the
equation:
2
P P
= 2
t x

(10)

This is the pressure equation for a 1D system where 0 x L, where L is the length
of the system. We can visualise this physically - much as we did in Chapter 5, Section
2.1 - using Figure 4. After the system is held constant at P = Po, the inlet pressure is
raised (at x = 0) instantly to P = Pin while the outlet pressure is held at Pout = Po.
Institute of Petroleum Engineering, Heriot-Watt University

Pin
t = t1

t = t2
Pout = Po

Po
t=0
0

These pressures, Pin and Pout, represent the constant pressure boundary conditions
(Mathematically these are sometimes called Dirichlet boundary conditions). In other
words, these are set, as if by experiment, and the system between 0 < x < L must sort
itself out or "respond" by simply obeying equation 10 above. We now approach
this problem using finite differences as follows:
discretise the x-direction by dividing it into a numerical grid of size x;
choose a time step, t;
use the following notation

Pin
Pinn +1
P
Piin
n +1
P
Pin +1
i

time level n = 0, 1, 2 ...


x-grid block label, i = 1, 2, 3 ... NX (at x = L)
current (known) P at time level
new (unknown) P at time level

fix the boundary conditions which, in this case, are as follows (see Figure
4):
P1 = P in and PNX = Pout = Po which are fixed for all t.
apply finite differences to equation 10 using the above notation to obtain:

P n +1 Pi n
P
i
t i
t

(11)

and

2 P
Pi ??+1 + Pi ??1 2 Pi ??
2
x i
x 2

(12)

However, an issue arises in equation 12 above as shown by the question marks on


the spatial derivative time levels. It is simply: which time level should we take for
the spatial derivative terms in equation 12? This is important and we will return to
this matter soon. However, for the moment, let us take these spatial derivatives at
time level n (the known level) since this will turn out to be the simplest thing we
can do. Thus, we obtain:

10

Figure 4
Physical picture of pressure
propagation in a 1D (compressible) system described
by. P 2 P

= 2
t x

Numerical Methods in Reservoir Simulation

2 P
Pi n+1 + Pi n1 2 Pi n
2
x i
x 2

(13)

Equating the numerical finite difference approximations of each of the above derivatives
as required by the original PDE, equation 10 (i.e. equating the expressions in equations
11 and 13) gives:

Pi n +1 Pi n Pi n+1 + Pi n1 2 Pi n

t
x

(14)

which easily rearranges to obtain an explicit expression for, Pi n +1, the only
unknown in the above equation:

Pi n +1 = Pi n +

t
( Pi n+1 + Pi n1 2 Pi n )
x 2

(15)

In words, we can interpret this above algorithm as saying:


New (n+1 level) value of Pi = Old (n level) value of Pi + a correction term
Equation 15 gives the algorithm for propagating the solution of the PDE forward in
time from the given set of initial conditions.
We now consider how to set the initial conditions. The initial conditions are the values
of all the P0i (i = 1, 2, 3 ... , NX) at t = 0. From Figure 4, these are clearly:

P10
P10
P1 = 1 for all time (boundary condition)
Pi 00
Pi 0 = 0 at time t = 0 for 2 < i < NX-1
Pi
0
= Po for all time (boundary condition)
PNX
0
PNX
0
PNX
Let us now
apply the above algorithm to the solution of equation 10. For this
problem, suppose we take the following data:
0 x 1.0
x = 0.1 => implies 11 grid points, P1 , P2 , P3, .... , P11 (NX = 11).
t = 0.001
n +1

The problem is then to calculate the solution, P(x,t), - that is Pi for all i (at all
grid points) and all future times up to some final time (possibly when the equation
comes to a steady-state as will happen in this example). This can be done by filling
in the solution chart Table 1 - see Exercise 2 below.

Institute of Petroleum Engineering, Heriot-Watt University

11

Grid Blocks
Time
x=
t

0.1 100

i=
P0
i

(IC)
P100
i

0.2 200

0.0
0.1
(BC)
1
2

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

10

1.0
(BC)
11

P200
i

0.3 300

P300
i

0.4 400

P400
i

0.5 500

P500
i

0.6 600

P600

0.7 700

P700
i

0.8 800 P800


i

0.9 900

P900
i

1 1000

P1000
i

Note:

BC = Boundary Conditions - these points are fixed;


IC = Initial Conditions, i.e. values of P(x, t = 0) for all i at t= 0

EXERCISE 2.

Fill in the above table using the algorithm (where (t/x2)=0.1):

Pi n +1 = Pi n +

t
P n + Pi n1 2 Pi n )
2 ( i +1
x

Hint: make up a spread sheet as above and set the first unknown block (shown grey
shaded in table above) with the above formula. Copy this and paste it into all of the
cells in the entire unknown area (surrounded by bold border above).
Answer: If you get stuck, look at spreadsheet CHAP6Ex2.xls on the disk.

12

Table 1.
Solution chart for the solution of the simple pressure
equation

Numerical Methods in Reservoir Simulation

The assumption we made in equation 12 was that the spatial derivative was taken at
the n (known) time level. This allowed us to develop an explicit formula for Pi n +1
(for all i). This method is therefore known as an explicit finite difference method.
We can learn about some interesting and useful features of the explicit method
which is encapsulated in equation 15 by simply experimenting with the spreadsheet
CHAP6Ex2.xls. Three features of this method can be illustrated by numerical
experiment as follows:
(i)

the effect of time step size, t;

(ii) the effect of refining the spatial grid size, x (or number of grid cells, NX);
(iii) the effect of running the calculation to steady-state as t (in practice,
until the numerically computed solution stops changing).
It is best if you do these yourself by modifying the spreadsheet (CHAP6Ex2.xls).

EXERCISE 3.

Experiment with the spreadsheet in CHAP6Ex2.xls to examine the effects of the


three quantities above - t, x (or NX) and the solution as t .

HINTS for Exercise 6.3:

Sensitivity to t: When you make t too big, the predicted numerical solution of
the PDE goes badly wrong - indeed negative pressures can occur which is physically
impossible. In fact, the solution has become unstable for the larger time steps.
This means that this explicit numerical method does have some time step limitations
which we must be careful of.

Sensitivity to x: the effects of grid refinement are that, as the grid blocks get
smaller (i.e. x 0 or NX ). The answer should get more accurate although,
to make this happen, you need to reduce the time step as well.

Behaviour as t : Finally, considering the long-time behavior of the


solution of the PDE, you should find that P(x, t ) tends to a straight line. Some
other intermediate pressure profiles can also be plotted using CHAP6Ex2.xls. In
fact, a little bit of analysis shows us that this is quite expected. As t , then if
steady-state is reached, then:

2 P
P
= 0, which implies => 2 = 0
t
x
But the curve with a zero second derivative (i.e. a first derivative which is constant)
is a straight line. Therefore, this result is physically reasonable and our numerical
model appears to be behaving correctly.
Institute of Petroleum Engineering, Heriot-Watt University

13

Another way to represent this explicit finite difference solution to the PDE is shown
in Figure 5, where we indicate the time levels for the solution of the equations and
we also show the dependency of the unknown pressure ( Pi n +1).
n

P i-1

Time

Pi

P i+1
Time level n

i-1

i+1

n+1

Pi

n+1

P i+1

t
Time level n + 1

P i-1

n+1

3.2 Implicit Finite Difference Approximation of the Linear Pressure


Equation

We now return to the original finite difference equation 12, where we had to

2 P
2
make a choice of time level for the spatial derivative, x . We now examine
the consequences of taking this derivative at the (n+1) time level - this is the
unknown time level. The finite difference equation for this case is as follows:
The time derivative is the same, i.e.

P n +1 Pi n
P
i
t i
t

(16)

but the spatial derivative now becomes:

2 P
Pi n+1+1 + Pi n1+1 2 Pi n +1
2
x i
x 2

(17)

As before, we now equate the numerical finite difference approximations of each of the
above derivatives (as required by the original PDE, equation 10) to obtain:

Pi n +1 Pi n Pi n+1+1 + Pi n1+1 2 Pi n +1

t
x

(18)

n +1
This equation shouldPibe
compared with equation 14. In the previous case, this could
n
easily be rearranged into
Pi +1 an explicit equation for Pi n +1 (equation 15). However,
equation 18 above cannot be rearranged to give a simple expression for the pressure
n +1
n +1
Pi n+1+1 there now appear to be three unknowns at
at the new time step P
(n+1),
.and
Indeed,
i 1 , Pi
1
n +1
and
Pi n1+1 , Ptoi n +be
time level (n+1), viz. Pi n1+1 , Pi n +1 and Pi n+1+1 . This appears
a bitPiof
+1 a paradox:
how do we find three unknowns from a single equation (equation 18 above)? The
answer is not really too difficult: Basically, we have an equation - like equation 18
- at every grid point. We will show below that this leads to a set of linear equations
where we have exactly the same number of unknowns as we have linear equations.

14

Figure 5.
Schematic of the explicit
finite difference algorithm
for solving the simple
pressure equation (a PDE).

Numerical Methods in Reservoir Simulation

Therefore, if these equations are all linearly independent, then we can solve them
numerically. This is a bit more trouble than our earlier simple explicit finite difference
method. Because, we do not get an explicit expression for our unknowns - instead
we get an implicit set of equations - this method is known as an implicit finite
difference method.
To see where these linear equations come from, rearrange equation 18 above
to obtain:

x 2 n
x 2 n +1
n +1
n
+
1
Pi n1+1 2 +

P
P
+
=
i

P
i +1

t i
t Pi

(19)

where all the unknowns ( Pi n1+1 , Pi n +1 and Pi n+1+1 ) are on the LHS of the equation
and the term on the RHS is known, since it is at the old time level, n. We
can write this equation as follows:

ai 1 Pi n1+1 + ai Pi n +1 + ai +1 Pi n+1+1 = bi

(20)

x 2 n
x 2
and where ai 1 = 1; ai = 2 +
;
1
and

a
b
=
=
i +1

P
i

t i
t

are all constants. The ai do not change throughout the calculation but the quantity bi is
updated at each time step as the newly calculated Pn+1 is set to the Pn for the next time
step. We can see how this works for the 5 grid block system in Figure 6 below:
1

P1

3
n

Time level n

P2

4
n

P3

P4

P5

x
t

Time level n=1

Figure 6.
Simple example of a 5 grid
block system showing how
the implicit finite difference
scheme is applied

P n+1
1

P n+1
2

P n+1
3

Boundary
condition
(fixed)

P n+1
4

P n+1
5
Boundary
condition
(fixed)

At each grid point, then (a) we know the value from the boundary condition (i =
1 and i = 5), (b) it uses a boundary condition (i=2 and i=4) or (c) it is specified
completely by equation 20 (only i = 3, in this case - but it would be most points
for a large number of grid points).
Consider each grid point in turn as follows:
i = 1 a boundary; therefore P1 is fixed, say as P1

Institute of Petroleum Engineering, Heriot-Watt University

15

P1

+ a2 P2n +1

P3n +1 =

=>

a2 P2n +1

P3n +1 =

i=2

i=3

i=4

P2n +1

+ a3 P3n +1

P3n +1

+ a4 P4n +1

=>

P3n +1 +

x 2 n

P
t 2
x 2 n

P P1
t 2

b2 =

P4n +1 =

P5 =

a4 P4n +1

b3 =

x 2 n

P
t 3

x 2 n

P
t 4
x 2 n

P P5
t 4

b4 =

iP=5 5 a boundary; therefore P5 is fixed, say as P5


Therefore there are only three unknowns in the above set of linear equations
1
n +1
( P2n +as
,P
and P4n +1 )
follows:
( P2n +1 , P3n +1 and P4n +1 ) which can be summarised
3

i = 2:

a2 P2n +1

n +1
2

i = 3:

P3n +1

n +1
3 3

+ aP

n +1
3

i = 4:

n +1
4

n +1
4 4

+ aP

x 2 n

P
t 2

P1

x 2 n

P
t 3
x 2 n

P
t 4

P5

(21)
Note that the set of linear equations above can be represented as a simple
matrix equation as follows:

a2

1
0

1
a3
1

1
a4

P2n +1
n +1
P3
n +1
P4

x 2 n
P2

t
x 2
n

P3

2
x P n
t 4

P1

P5

(22)

16

Numerical Methods in Reservoir Simulation

The structure of this matrix equation


clearer when there are more equations
P1 and Pis
12
involved. For example, it is quite easy to show that, if we take 12 grid points instead
of the 5 above, we obtain 10 equations (using the two fixed boundary conditions,
P1 and P12 ) for the quantities, P2n +1 , P3n +1 , P4n +1 ....P11n +1, of the form:
n +1
2a2

P
1

0
0

0
0

0
0

n +1
13

, P

n +1
4 0

,0P

n +1
110

....P

a3

a4

a5

a6

a7

a8

a9

a10

0
0

0
0

0
0

1
a11

P2n +1
n +1
P3
n +1
P4
P n +1
5
P6n +1
n +1
P7
n +1
P8
P n +1
9
P10n +1
n +1
P11

( x 2

( x 2

( x 2

( x 2

( x 2
=
( x 2

( x 2

( x 2

( x 2

( x 2

t ) P2n P1

t ) P3n

t ) P4n

t ) P5

t ) P6n

t ) P7n

t ) P8

t ) P9n

t ) P10n

t ) P11n P12

(23)
And 20 grid points in 1D would lead to the following set of 18 equations:

a2
1

0
0

0
0

0
0

0
0

0
0

0
0

1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
a3 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

1 a4 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0 1 a5 1 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 1 a6 1 0 0 0 0 0 0 0 0 0 0 0 0

0 0 0 1 a7 1 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 1 a8 1 0 0 0 0 0 0 0 0 0 0

0 0 0 0 0 1 a9 1 0 0 0 0 0 0 0 0 0

0 0 0 0 0 0 1 a10 1 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1 a11 1 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0 1 a12 1 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 1 a13 1 0 0 0 0 0

0 0 0 0 0 0 0 0 0 0 1 a14 1 0 0 0 0

0 0 0 0 0 0 0 0 0 0 0 1 a15 1 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 1 a16 1 0 0

0 0 0 0 0 0 0 0 0 0 0 0 0 1 a17 1 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 a18 1

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 a19

P2n +1
n +1
P3
n +1
P4
P n +1
5
P6n +1
n +1
P7
n +1
P8
P n +1
9
P10n +1
n +1
P11
n +1
P12
P n +1
13
P n +1
14
P15n +1
n +1
P16
n +1
P17
P n +1
18
P19n +1

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2
=
( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

( x 2

x 2
(

t ) P2n P1

t ) P3n

t ) P4

t ) P5n

t ) P6

t ) P7n

t ) P8n

t ) P9n

t ) P10n

t ) P11n

t ) P12n

t ) P13n

t ) P14n

t ) P15n

t ) P16n

t ) P17n

t ) P18n

n
t ) P19 P20

(24)
Institute of Petroleum Engineering, Heriot-Watt University

17

For this simple 1D PDE, it is clear that the matrix arising from our implicit finite
difference method has the following properties:
(i) It is tridiagonal - that is, it has a maximum of three non-zero elements in any
row and these are symmetric around the central diagonal;
(ii) It is very sparse - that is, most of the elements are zero. In an MxM matrix,
there are only 3M non-zero terms but M2 actual elements. If M = 100, then the
matrix is only (300/1002)x100% = 3% filled with non-zero terms.
As it happens, a very simple computational procedure, called the Thomas algorithm,
can be used to solve tridiagonal systems very quickly. The FORTRAN code
for this is shown for interest in Figure 7. Note that it is very compact and
quite simple in structure. You are not expected to know this or to necessarily
understand how this algorithm works.

18

Numerical Methods in Reservoir Simulation

THE THOMAS ALGORITHM

Figure 7.
The Thomas algorithm for
the solution of tridiagonal
matrix systems.

subroutine thomas
c Thomas algorithm for tridiagonal systems
+
(N ! matrix size
+
,a ! diagonal
+
,b ! super-diagonal
+
,c ! sub-diagonal
+
,s ! rhs
+
,x ! solution
+
)
c---------------------------------------c This program accompanies the book:
c C. Pozrikidis; Numerical Computation in Science and Engineering
c Oxford University Press, 1998
c-----------------------------------------------c Coefficient matrix:
c | a1 b1 0 0 ... 0 0 0 |
c | c2 a2 b2 0 ... 0 0 0 |
c | 0 c3 a3 b3 ... 0 0 0 |
c | .............................. |
c | 0 0 0 0 ... cn-1 an-1 bn-1 |
c | 0 0 0 0 ... 0 cn an |
c-----------------------------------------Implicit Double Precision (a-h,o-z)
Dimension a(200),b(200),c(200),s(200),x(200)
Dimension d(200),y(200)
Parameter (tol=0.000000001) this is a measure of how close to coverage we are
c prepare
Na = N-1
c reduction to upper bidiagonal
d(1) = b(1)/a(1)
y(1) = s(1)/a(1)
DO i=1,Na
i1 = i+1
Den = a(i1)-c(i1)*d(i)
d(i1) = b(i1)/Den
y(i1) =(s(i1)-c(i1)*y(i))/Den
End DO
c Back substitution
x(N) = y(N)
DO i=Na,1,-1
x(i)= y(i)-d(i)*x(i+1)
End DO
c Verification and alarm
Res = s(1)-a(1)*x(1)-b(1)*x(2)
If(abs(Res).gt.tol) write (6,*) thomas: alarm
DO i=2,Na
Res = s(i)-c(i)*x(i-1)-a(i)*x(i)-b(i)*x(i+1)
If(abs(Res).gt.tol) write (6,*) thomas: alarm
End DO
Res = s(N)-c(N)*x(N-1)-a(N)*x(N)
If(abs(Res).gt.tol) write (6,*) thomas: alarm
c Done
100 Format (1x,f15.10)
Return
End
Institute of Petroleum Engineering, Heriot-Watt University

19

Finally, we note that the implicit finite difference method can be viewed as shown
in Figure 8 below. The choice of the (n+1) time level for the spatial discretisation
leads to a set of linear equations which are somewhat more difficult to solve than the
simple algebraic equation that arises in the explicit method (equation 15). However,
there are exactly the same number of linear equations as there are unknowns and so
it is possible to solve these.
n

P i-1

Time

Pi

P i+1
Time level n

i-1

i+1

n+1
P i-1

n+1
Pi

n+1
P i+1

t
Time level n + 1

Figure 8.
Schematic of the implicit
finite difference algorithm
for solving the simple pressure PDE

3.3 Implicit Finite Difference Approximation of the 2D Pressure


Equation

Before we go on to discuss how to solve the sets of linear equations that arise in the
finite difference approximation of PDEs arising in reservoir simulation, we will first
consider the discretisation of the 2D single-phase pressure equation. This raises
some additional important issues which occur when we try to solve more complicated
systems, as follows:
(i)

the complication of the more connected grid block system in a 2D


domain;

(ii) the possibility of heterogeneity in the permeability field which leads us to


the matter of how to take average properties in grid-to-grid flows between
blocks of different permeability (dealt with in Chapter 4);
(iii) the issue of non-linearity for a compressible system e.g. c(P) is clearly a
function of P(x,t), which is the unknown.

3.3.1 Discretisation of the 2D Pressure Equation

We take as the basic pressure equation for single phase slightly compressible flow,
the following (see the solution for exercise 2 at the end of Chapter 5):

c P P P
kx +
=
k y
k t x x y y
k x and k y
c
k

(25)

c
k x and k y

(a
k simplified form of equation 39, Chapter 5) where the term k is a constant
which we will denote as below, k is an average permeability of the entire
reservoir and k x and k y are the local permeabilities normalised by k ; i.e.
k non-linearities for
k x = k x / k and k y = ky / k. Note that we have avoided the
the moment (the quantities c(P), and usually depend on pressure). However, the

20

k x = k x / k and k y = ky / k.

k x = k x / k and k y = ky / k.

Numerical Methods in Reservoir Simulation

system may be anisotropic (kx ky) and heterogeneous (the permeability may vary
from grid block to grid block).
Equation 25 above can be discretised in a similar way to that applied to the 1D
linear pressure PDE discussed above. However, we will need to be quite clear
about our notation in 2D and, for this purpose, we refer to Figure 9. This shows the
discretisation grid for the above PDE - note that this is essentially the opposite of
what we did when we derived the equation in the first place! In Chapter 5, we used
a control volume (or grid block) to express the mass conservation and then inserted
Darcys law for the block to block flows; we then took limits as x, y and t
0. Here, we are starting with the PDE and going back to the local conservation of
flows and introducing finite size x and y.

i, j + 1

(j + 1/2)
y

i - 1, j

i, j

i + 1, j

(j - 1/2)

Figure 9.
Discretisation and notation
for the 2D pressure
equation.

i, j - 1

y (j)

(i x (i)

1/ )
2

(i + 1/2)

Discretising the following equation using the above notation

P P P
= kx + ky
t x x y y
we obtain:
P
P
k x
k x x
P P

i +1/ 2 x i 1/ 2

t
x
n +1

(26)
P
P
k y
k y
y j +1/ 2 y j 1/ 2
y

(27)
where the (i 1/2) and (j 1/2) subscripts refer to quantities at the boundaries as
shown in Figure 9. We can now expand these boundary flows as follows:

Institute of Petroleum Engineering, Heriot-Watt University

21

P n +1 Pi ,nj
i, j
= kx
t

( )

Pi n+1+,1j Pi ,nj+1

kx
i +1 / 2
x 2

( )

+ k y

( )

Pi ,nj+1 Pi n1+,1j

i 1 / 2
x 2

Pi ,nj++11 Pi ,nj+1

ky
j +1 / 2
y 2

( )

Pi ,nj+1 Pi ,nj+11

j 1 / 2
y 2

(28)

( )

( )

( )

( )

where the quantities k x i +1/ 2 , k x i 1/ 2 , k y j +1/ 2 and k y j 1/ 2 are some type of


average permeabilities between the two neighbouring grid blocks - as discussed in
Chapter 4. Note also we have chosen the spatial discretisation terms at the new (n+1)
time level making the this an implicit finite difference scheme.
We can rearrange equation 28 above by taking all the unknown terms (at n+1) to the
LHS and the known terms (at time level n) to the RHS. This gives the following:

( )

( )

( )

( )

( )

( )

( )

( )

k x

k x

k y
k y
k x
x
n +1
j 1 / 2
j +1 / 2
n +1
n +1
i 1 / 2
i +1 / 2
i 1 / 2
i +1 / 2

Pi , j

P
+
+
+
+
i 1, j
i +1, j
2
t
x 2
y 2
y 2
x 2
x 2
x

k x

k y

n
j 1 / 2
j +1 / 2
n +1

Pi ,nj++11 =

Pi , j
i , j 1
2
2
t
y
y

(29)
Since the coefficients in equation 29 above are constants, then this defines a set of
linear equations similar to those found in 1D. However, here we have up to five
non-zero terms per grid block to deal with rather than the three we found for the
1D system. The matrix which arises in this 2D case is known as a pentadiagonal
matrix. This set of linear equations can be written as follows:

ai 1, j .Pi n1+,1j + ai +1, j .Pi n+1+,1j + ai , j .Pi ,nj+1 + ai , j 1 .Pi ,nj+11 + ai , j +1 .Pi ,nj++11 = bi , j
(30)
where the constant coefficients, ai 1, j , ai +1, j , ai , j , ai , j 1 and ai , j +1 , are given by the
coefficients in equation 29; bi, j is also a (know) constant.

3.3.2 Numbering Schemes in Solving the 2D Pressure Equation

It is quite convenient to label the pressures as Pi, j when we are working out the
discretisation of the equations, but this is not helpful when we are arranging the
linear equations. Here, it is useful first to consider the numbering scheme for the
2D system that allows us to dispense with the (i,j) subscripting in equations 29 or 30
above. The structure of the A - matrix in equation 30 can be made clearer by working
out a specific 2D example as shown in Figure 10.

22

Numerical Methods in Reservoir Simulation

j=5
j=4
j=3
j=2
j=1

Figure 10.
Numbering system for 2D
grid conversion from (i,j)
m counter.

m = 17
m = 13
m=9
m=5
m=1
i=1

m = 18
m = 14
m = 10
m=6
m=2
i=2

m = 19
m = 15
m = 11
m=7
m=3
i=3

m = 20
m = 16
m = 12
m=8
m=4
i=4

Notation:
NX
= maximum number of grid blocks in x - direction, i = NX;
NY
= maximum number of grid blocks in y - direction, j = NY
m
= grid block number in the natural ordering scheme shown
m
= (j - 1).NX + i
e.g. for i = 3, j = 4 and NX = 4, m = (4 - 1).4 + 3 = 15 (as above)

In the m-notation shown in Figure 10 (m = (j-1)NX +i), the reordered equations 30


become the following:

a m1 .Pmn+11 + a m+1 .Pmn++11 + a m .Pmn +1 + a m NX .Pmn+1NX + a m+ NX .Pmn++1NX = bm


(31)
where the a m are the reordered coefficients where the subscript is calculated from
the m-formula in Figure 10. For example:

ai , j 1 a ( j 11) NX +i =( j 1) NX +i NX a m NX

(32)

ai , j +1 a ( j +11) NX +i = ( j 1) NX +i + NX a m+ NX

(33)

Note that, when we apply the above equation numbering scheme to the example in
Figure 10 (NX = 4, NY = 5 and therefore, 1 m 20), certain neighbours are
missing since a block is at the boundary (or in a corner where two neighbours are
missing). For example, for block (i = 3; j = 3), that is block m = 9, the j-1 block is
not there. Therefore, the coefficient Am-1 =0 in this case. This is best seen by writing
out the structure of the 20 x 20 A-matrix by referring to Figure 10; the A-matrix
structure is shown in Figure 11.
Note that the A-matrix structure in Figure 11 is sparse and has a maximum of five
non-zero coefficients in a given row - it is a pentadiagonal matrix.
All implicit methods for discretising the pressure equation lead to sets of linear
equations. These have the general matrix form:

A.x = b

(34)

where A is a matrix like the examples shown above, x is the column vector of unknowns
(like the pressures) and b is a column vector of the RHSs. This is just like equation
31 but it is in shorthand form. We will discuss methods for solving these equations
later in this chapter. For the meantime, we will just assume that it can be solved. We
next consider when the PDEs describing a phenomenon are non-linear PDEs.

Institute of Petroleum Engineering, Heriot-Watt University

23

m 1 2
x x
1
x x
2
x
3
4
5 x
x
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

3
x
x
x

5
x

9 10 11 12 13 14 15 16 17 18 19 20

x
x

x
x

x
x

x
x

x
x

x
x
x
x

x
x
x

x
x
x
x

x
x

x
x
x

x
x

x
x
x

x
x

x
x
x

x
x

x
x
x

x
x

x
x
x

x
x

x
x
x

x
x

x
x
x

x
x
x

x
x

m - ordering scheme used in figure 10 - shown here for reference.


j=5
j=4
j=3
j=2
j=1

m = 17
m = 13
m=9
m=5
m=1
i=1

m = 18
m = 14
m = 10
m=6
m=2
i=2

m = 19
m = 15
m = 11
m=7
m=3
i=3

Figure 11.
Structure of the A-matrix
for the m-ordering scheme
in the table shown below for
reference.

m = 20
m = 16
m = 12
m=8
m=4
i=4

3.4 Implicit Finite Difference Approximation of Non-linear Pressure


Equations

We have noted previously that using numerical methods are usually the only way
which we can solve non-linear PDEs of the type that arise in reservoir simulation.
We will use the single-phase 2D equation for the flow of compressible fluids (and
rocks) as the example for discussing the numerical solution of non-linear PDEs.
Equation 35 below was derived in Chapter 5 and has the form:

P kx P ky P
c( P) =
+
t x x y y

(35)

where the non-linearities arise in this equation due to the dependence of the generalised
compressibility, c(P), the density, (P) and the viscosity, (P), on pressure, P(x,t). It
is the pressure that is the main unknown in this equation and, hence, if these quantities
depend on it, then they introduce difficulties into the equations. In fact, we will see
shortly that the equations that arise are not linear equations - they are non-linear
algebraic equations.

24

Numerical Methods in Reservoir Simulation

Proceeding as before, we can discretise the above equation as follows:


ky P
k P
kx P
kx P
y

P n +1 P n x i +1/ 2 x i 1/ 2 y j +1/ 2 y j 1/ 2
c( P)
+
=

t
x
y

(36)
where we have not yet decided on the time level of the non-linearities (i.e. the time
level of the pressure, Pn or Pn+1, at which we evaluate c, and ). The most difficult
case would be if these were set at the unknown time level Pn+1 and this is what we
will do as follows:

k n +1 P

x
n +1
n
x

P
P

i +1 / 2
c( P n +1 )
=

t
x

k n +1 P
x
x i 1/ 2

ky n +1 P


y j +1/ 2
+
y

ky n +1 P


y j 1/ 2
(37)

P
P
and terms gives:
x
y

Now expanding up the

k n +1 Pi n+1+,1j Pi ,nj+1 k n +1 Pi ,nj+1 Pi n1+,1j


x
x

n +1
n

x
x

P
P

i +1 / 2
i 1 / 2
i, j
i, j
n +1

c( P )
=

t
x

ky n +1 Pi ,nj++11 Pi ,nj+1 ky n +1 Pi ,n1+1 Pi ,nj+11

y
y


j +1/ 2
j 1 / 2
+
y

(38)

which can be simplified to the following:


n +1

c( P

n +1

n +1

P n +1 Pi ,nj kx
k
Pi n+1+,1j Pi ,nj+1 x 2
Pi ,nj+1 Pi n1+,1j
) i , j
=

2
t
.x i 1/ 2

.x i +1/ 2

n +1

k
+ y 2
Pi ,nj++11 Pi ,nj+1
.y j +1/ 2

n +1

Pi ,nj+1 Pi ,nj+11
k
y 2

y
.y j 1/ 2

(39)

Institute of Petroleum Engineering, Heriot-Watt University

25

This can be treated as before to gather together similar unknown P-terms as


follows:
n +1

n +1

k
k
x 2
.Pi n1+,1j y 2
.Pi ,nj+11
.x i 1/ 2
.y j 1/ 2
n +1
c( P n +1 ) k n +1 k n +1 ky n +1
n +1
ky
x
x
+
+
+
+
+
.Pi , j

2
2
2
2
t .x i +1/ 2 .x i 1/ 2 .y j +1/ 2 .y j 1/ 2

n +1

n +1

c( P n +1 ).Pi ,nj
k
k
.Pi ,nj++11 =
x 2
.Pi n+1+,1j y 2

t
.y j +1/ 2
.x i +1/ 2

(40)
As before, we can write this in a compact form as follows:

in+11, j .Pi n1+,1j + in++11, j .Pi n+1+,1j + in, +j 1 .Pi,nj+1 + in, +j 11 .Pi,nj+11 + in, +j +11 .Pi,nj++11 = in, +j 1
(41)
where the elements of theA-matrix (now denoted in+11, j , in, +j 11 , in, +j 1 , in++11, j and in, +j +11 )
are not constants since they depend on the (unknown) value of Pn+1.
How do we go about solving the non-linear set of algebraic equations in 39 or 40
above? It turns out that we have two choices in tackling this more difficult problem
as follows:
(i) We can actually use a numerical equation solver which is specifically designed
to solve more difficult non-linear problems. An example of this type of approach
is in using the Newton-Raphson method. We will return to this method later (once
we have seen how to solve linear equations).
(i) We can choose to apply a more pragmatic algorithms such as the following:
(a) Although our -terms in equation 41 are strictly at time level (n+1), simply
take them at time level n, as a first guess. So we approximate equation 41 by the
following first guess:

in1, j .Pi n1+,1j + in+1, j .Pi n+1+,1j + in, j .Pi,nj+1 + in, j 1 .Pi,nj+11 + in, j +1 .Pi,nj++11 = in, j
n
n
n
in1, j , in+1, j , in, j , i(42)
, j 1 , i , j +1 and i , j

Pi ,nj+1

n +1

Pi , at
j the
where the quantities in1, j , in+1, j , in, j , in, j 1 , in, j +1 and in, j are evaluated
n
(known) time level
Pi ,nj+1
n - i.e. at values of pressure = Pi, j .
i, j

n +1

Pi , 42
j
(b) Solve the now linear
above to obtain a first estimate - or ia, j first
Pi ,nj equations

iteration - of the Pi ,nj+1at each grid point. We will use the following notation

i, j

where is the iteration counter and i , j is the value of the a-coefficient at the Pi ,nj+1
value after iterations; that is:

[(

i, j = i , j Pi,nj+1

26

)]

(P )

n +1
i, j

(43)

(P )

n +1
i, j

Numerical Methods in Reservoir Simulation

Pi ,nj+1

(c) Use the latest iterated values of the i , j to solve the (now linear) equations to

go from

(P )

n +1
i, j

to Pi ,nj+1

+1

n +1

(d) Keep iterating the above scheme untilPit


i , j converges; i.e. the difference between
the sum of two successive iterated values of the pressure (Err.) is sufficiently small
(< Tol, which is an acceptably small value)

Err =

(P )

n +1 +1
i, j

Pi ,nj+1

all i , j

(44)

Stop if Err < Tol. Otherwise continue through steps above.


The algorithm outlined in (ii) above for solving the non-linear pressure equation
is represented in Figure 12.
Set the iteration counter = 0

Calculate the i1, j , i+1, j , i, j , i, j 1 , i, j +1 and i, j at

( )

n
values of pressure = Pi , j

[(

i, j = i , j Pin, j+1

= +1

)]

where = current iteration number.

Solve the set of linear equations:


i1, j .Pin+11, j + i+1, j .Pin++11, j + i, j .Pin, j+1 + i, j 1 .Pin, j+11 + i, j +1 .Pin, j++11 = i, j

to obtain Pin, j+1 = Pin, j+1

+1

Calculate
Err =

(P )

n +1 +1
i, j

all i , j

No - continue iterations

Figure 12.
Algorithm for the numerical
solution of the non-linear
2D pressure equations:

Pi ,nj+1

Err < Tol ?

Yes (the method has converged)

Stop

Institute of Petroleum Engineering, Heriot-Watt University

27

4
APPLICATION OF FINITE DIFFERENCES TO TWO-PHASE
FLOW
4.1 Discretisation of the Two-Phase Pressure and Saturation
Equations

We now consider how finite difference methods are applied to the two-phase flow
equations. We have seen how these equations were derived in Chapter 5. Recall that,
in two-phase flow, we have two coupled equations to solve - a pressure equation and
a saturation equation. For example, we may solve for the oil pressure, Po(x,t) and the
oil saturation, So (x,t); we would then find the water saturation and water pressure by
using the constraint equations, So+Sw = 1 and the capillary pressure relation, Pc(Sw)
= Po - Pw, respectively.
From Chapter 5, we use the highly simplified form of the 1D pressure and saturation
equations, where we choose to solve for P(x,t) and So (x,t) as shown in Chapter
5, equations 81 and 82. Note that we have taken zero capillary pressure (therefore
P = Po = Pw) and zero gravity which gives:
PRESSURE EQUATION

SATURATION EQUATION


P
T ( So ) = 0

x
x
S
P
o = o ( So )
t x
x

(45)

(46)

(Equations 45 and 46 are the same as equations 81 and 82 from Chapter 5,


respectively).
(S )
T

The quantity T ( So ) is the total mobility and is the sum of the oil and water mobilities;

T ( So ) = o ( So ) + w ( So ) . The above two equations are clearly coupled together since

the oil saturation appears in the non-linear coefficient of the pressure equation. Likewise,
= the
T ( So ) in
(So ) + term
(Soon) the RHS of the saturation equation.
the pressure appears
o flow
w
We can now apply finite differences to each of these equations in the same
way as discussed above to obtain, for the pressure equation (see notation in
Figure 8):

P
P
T ( So )
T ( So ) x
x i 1/ 2

i +1 / 2
=0
x

(47)

which can be expanded further to give:

Pi n+1+1 Pi n +1
Pi n +1 Pi n1+1

S
(
)
(
)

T o
i +1/ 2 T o
i 1 / 2
x
x

=0
x
(48)

28

Numerical Methods in Reservoir Simulation

( (S ))
T

( (S ))
T

n +1
and STo(So )

i +1 / 2

i +1 / 2

and T (So )

i 1 / 2

i 1 / 2

6
)

In equation 48 above, we must now specify the time


non-linear
Tlevel
T (Somobility
(So )of the and
)

( ( ) )

and
T (Son +T1()So )

terms,
Son +1 T (So )

i +1 / 2

i 1 / 2

; we must choose whether we take these terms


i +1 / 2
i +1 / 2i 1 / 2
at time n or at time level (n+1)? Again, we go straight to the most difficult case by
n +1
defining these terms at the later time level i.e. at So . These terms are denoted as,

( (S S )) and ( (S )) , in equation 48 above to obtain:


( ( S ))
( ((S (S)) ))

S
)
(

(
)
( (S )) x ( P P ) x ( P P ) = 0

( (S ))
(49)
( (S ))
T

n +1
n +1
o
o

n +1
n +1 o
T
i +1 / 2
o
n +1 T
i +21 / 2
o
i 1 / 2

i +1 / 2

n +1
o

n +1
o

i 1 / 2

n +1
i +1

n +1
nT+1 o
i +1 / 2
o
n +1
i 1 / 2
i
2

n +1

n +1
o

n +1
i 1

i 1 / 2

i 1 / 2

The above equation can now be arranged into the usual order as follows:

( (S ))
T

n +1
o

x
n +1
i 1

i 1 / 2

, Pi

n +1

n +1
i 1

( Son +1 )
T

x 2

i 1 / 2

( (S ))
+
T

n +1
o

i +1 / 2

( S n +1 )
P n +1 + T o
i
x 2

i +1 / 2

Pi n+1+1 = 0

(50)

n +1
i +1

and P

This is a non-linear set of algebraic equations for the unknown pressures,


Pi n1+1 , Pi n +1 and Pi n+1+1 , but the coefficients depend on the - also unknown - saturations,
Son +1. Note that equation 50 represents one of a set of equations since there is one
at each grid point (and at the ends of the 1D system the values may be set by the
boundary
conditions).
Son +1
We now consider the discretisation of the saturation equation 46. Finite differences
may be applied to this equation as follows:

P
P
o ( So )
o ( So ) x
x i 1/ 2
S S
i +1 / 2

t
x
n +1
oi

n
oi

(51)

Again, we can expand the derivative terms at the (i+1/2) and (i-1/2) boundaries (Figure
9) and we can take the mobility terms at the (n+1) time level to obtain:

n +1

Soin+1 Soin o ( So )

t
x 2

i +1 / 2

(P

n +1
i +1

Pi

n +1

( Son +1 )
) o x 2

i 1 / 2

(P
i

n +1

n +1
i 1

(52)
The above non-linear algebraic set of equations can be written in various ways. Two
particularly useful ways to write the above equations are as follows:

Institute of Petroleum Engineering, Heriot-Watt University

29

Form A:
n +1
oi

n +1

t o ( So )
=S +

x 2

n
oi

i +1 / 2

(P

n +1
i +1

Pi

n +1

( Son +1 )
) o x 2

i 1 / 2

(P
i

n +1

(53)

or Form B:
n +1
oi

n +1
i 1

n +1

t o ( So )
S

x 2

n
oi

i +1 / 2

(P

n +1
i +1

Pi

n +1

( Son +1 )
) o x 2

i 1 / 2

n +1
n +1

P
P
(i
i 1 ) = 0

(54)

The reason for writing the above two forms of the saturation equation is that each is
useful, depending on how we intend to approach the solution of the coupled pressure
(equation 45) and saturation (equation 46) equations.
As with the case of the solution of the non-linear single phase pressure equation for
a compressible system, we have two strategies which we can use to solve the twophase equations above, as follows:
(i) We can actually use a numerical equation solver which is specifically designed to
solve more difficult non-linear problems; e.g. the Newton-Raphson method. We will
return to this method later (once we have seen how to solve linear equations).
(i) We can again choose to apply a more pragmatic algorithm and this is the subject
of section 4.2 below.

4.2 IMPES Strategy for Solving the Two-Phase Pressure and


Saturation Equations

The form of the discretised pressure and saturation equations which we will take are
as follows (equations 50 and 53):
Pressure:

( (S ))
T

n +1
o

i 1 / 2

n +1
i 1

( Son +1 )
T

x 2

i 1 / 2

( (S ))
+
T

n +1
o

i +1 / 2

( S n +1 )
P n +1 + T o
i
x 2

i +1 / 2

Pi n+1+1 = 0

(50)
Saturation:
Soin+1 + Soin

30

n +1

t o ( So )


x 2

i +1 / 2

( Son +1 )
n +1
n +1

P
P
( i +1 i ) o x 2

i 1 / 2

n +1
n +1

P
P
(i
i 1 )

(53)

Numerical Methods in Reservoir Simulation

Pressure equation 50 would be a set of linear equations if the coefficients were


known at the current time step (n), rather than being specified at the (n+1) - i.e. the
unknown - time level. However, we
could solve equation 50 above as if it were
Pi n1+1 , Pi n +1 and Pi n+1+1
n +1
1 n +nthe
n +(see
a linear system of equations by time-lagging
the coefficients
Pi n1+1as
, Pbefore
and
P +i1 1+1 and Pi n+1+1
P
i
i 1 , iP
algorithm in Figure 13). This would
give us a first guess (or first iteration) to find
Pi nn1++11, Pi nn++11and Pi n+n1++11
the unknowns i.e. the quantities Pin+11 , Pi and Pi +1 . The same problem exists
P
n +1
forPthe
equation
53, if wei had the first guess at the Pi n +1 (by thePprocedure
n +1saturation
n +1
n +1
i
,
P
and
P
i 1
i
i +1
+1
just mentioned), then we could use
latest values of pressure and still time lag
Pi nthese
n +1
the coefficients (the oil mobility Sterms)
Pni +1 and use the saturation equation 53 as if it
+1
o
n an
+1 explicit expression. This would give us an updated
Sonvalue
were
of Son +1 which
Pi
could then be used back in the pressure
Sonn++11 equation and the whole process could be

So n +1 previously.
iterated until convergence, as discussed
A flow chart
of this procedure
Pi 1 , Pi n +1 and Pi n+1+1 .n +1 n +1n +1 n +1n +1
n
+
1
hasSalready been outlined in Chapter 5 (Figure 9) but is elaborated
Figure
and
,here
.
P
Pi Pin
, Pi Pi +1 and
i 1
i 1
o
13. Note that by taking time-laggednvalues
ofn +the
+1
1 saturations,
n +1 the pressure equation
Pi n1 +1 , Pi n +1 and Pi +n1 +1 .
is linearlised and can then be solved
for the pressure for that iteration,
Pin+implicitly
11 , Pin +1 and Pin++11 .

( ) ( )
( () ) ( )
( ) (( ))
(( )) (( )) (( ))
( P ) , ( P ) and ( P )P obtained
( P ) , ( P ) and ( P ) . The saturations can then (be
) , ( P( Pexplicitly
) )and, ( P( P ) )and
using the latest pressures (i.e. the ( P ) , ( P ) and ( P ) ) and the most recent
( P ) , ( P ) and ( P )
iteration of the saturation (i.e. the ( S ) , ( S ) etc.). Therefore, this approach is
, ( PIMPES
) , (S(S Explicit
) )etc.
, ( S ) etc.
( P as) the
) and
( P ) which stands for IMplicit(Sin Pressure,
known
approximation,
S )it,converges
in Saturation. This can be iterated(until
although there are limitations in
((SS )) etc.
,
etc.
S
(
)
the size of time step, t, which can be taken. If t is too large, the IMPES method
may
(Sbecome
) , (Sunstable
) etc.and give unphysical results like the example in Exercise 3
n +1
i 1

n +1
i 1

n +1
i

above.

n +1

n +1

n +1
i +1

n +1
i +1

n +1
i +1

n +1
i 1

i 1

i +1

n +1
i n1 +1
ni +1
i

n +1
i n +1
in +1
i +1

n +1
i n +1
i

n +1
i +n1+1
i +1

n +1
i +n1 +1
i +1
n +1
i

Institute of Petroleum Engineering, Heriot-Watt University


n +1n +
1
i 1

n +1n +1
i i +1

n +1n +1
i +1i

n +1
i +1

31

Set the iteration counter to zero, = 0


Take current values of Son and P n as the level
iteration values: So and P

PRESSURE EQUN. Set the total mobilities T (So )i +1/ 2 and T (So )i 1/ 2
obtain the linearised pressure equation:

( (S ))

(So )

T (So )
T (So )
T
i 1 / 2
i +1 / 2 n +1
i +1 / 2
Pin+11
+
Pi +
Pin++11 = 0
2
2

x
x
x
x 2

Solve this linear implicit pressure equation for pressure at iteration, , to


obtain:
T

i 1 / 2

( Pin+11 ) , ( Pin +1 ) and ( Pin++11 )

SATURATION EQUN. Now update the saturation equation as if it were explicit by taking

(i) the oil mobility terms at the latest iteration, 0 (So )i 0 (So )i +1 , etc.

(ii) the latest values of the pressures just calculated implicity ( Pin+11 ) , ( Pin +1 ) and ( Pin++11 )
Update saturation explicity as follows:
Soin+1 = Soin +

t o (So ) i +1/ 2
(Pin++11 ) (Pin +1 )


x 2

to obtain latest iteration of saturations

(So )

i 1 / 2
o
(Pin +1 ) (Pin+11 )

x 2

(Sin +1 ) , (Sin++11 ) etc.

= +1

Calculate the error, Err., by comparing the change in pressure or saturations


(or both as here) at the current () and last (-1) iterations:
NX

Err. = ( Pin +1 ) ( Pin +1 )


i =1

NX

+ (Sin +1 ) (Sin +1 )

i =1

Note that the Err. term above would have some scaling factors weighing the
pressure and saturation terms because of the units of pressure.

No - continue iteration

32

Is Err. < Tol ?

Yes

Stop

Figure 13.
IMPES Algorithm for the
numerical solution of the
two-phase pressure and saturation equations (IMPES
IMplicit in Pressure,
Explicit in Saturation)

Numerical Methods in Reservoir Simulation

THE NUMERICAL SOLUTION OF LINEAR EQUATIONS

5.1 Introduction to Linear Equations

In all implicit finite difference methods, we end up with a set of linear equations to
solve. In fact, we have shown above that the equations involved - e.g. the discretised
pressure equation - may be a set of non-linear algebraic equations. However, these
can usually be linearised (say, by time lagging the coefficients as discussed above)
in order to obtain a set of linear equations. The solution to the original non-linear
equations may then be found by repeating or iterating through the cycle of solving
the linear equations. In the following section, we will address the issue of solving
the non-linear equations which arise in the discretised two-phase flow equations
directly.
A generalised set of linear equations may be written in full as follows:

a11 . x1 + a12 . x2 + a13 . x3 + a14 . x4 + a15 . x5 ... + a1n . xn = b1


a21 . x1 + a22 . x2 + a23 . x3 + a24 . x4 + a25 . x5 ... + a2 n . xn = b2
a31 . x1 + a32 . x2 + a33 . x3 + a34 . x4 + a35 . x5 ... + a3n . xn = b3
.....
.....

...
...

.....
...
an1 . x1 + an 2 . x2 + an 3 . x3 + an 4 . x4 + an 5 . x5 ... + ann . xn = bn

(55)

where the aij and bi are known constants and the xi (i = 1, 2, ... , n) are the unknown
quantities which we are trying to find. The xi in our pressure equation, for example,
would be the unknown pressures at the next time step.
The set of linear equations can be written in matrix form as follows:

a11

a21

a31

.....
.....

.....

an1

a12

a13

a14

a22

a23

a24

a32

a33

a34

an 2

an 3

an 4

a15 ... a1n x1 b1




a25 ... a2 n x2 b2


a35 ... a3n x3 = b3

... . .
... . .

... . .

an 5 ... ann xn bn

Institute of Petroleum Engineering, Heriot-Watt University

(56)

33

We can write the matrix A and the column vectors x and b as follows:

a11

a21

= a31

.....
.....

.....

an1

a12

a13

a14

a22

a23

a24

a32

a33

a34

an 2

an 3

an 4

a15 ... a1n

a25 ... a2 n

a35 ... a3n ;

...
...

...

an 5 ... ann

x1


x2


x = x3 ;
.
.

.
x
n

and

b1


b2


b = b3

.
.

.
b
n
(57)

and the set of linear equations can be written in very compact form as follows:

A.x = b

(58)

where A in an n x n square matrix and x and b are n x 1 column vectors.


A very simple example of a set of linear equations is given below:

4. x1 + 2. x2 + 1. x3 = 13
1. x1 + 3. x2 + 3. x3 = 14
2. x1 + 1. x2 + 2. x3 = 11

(59)

How do we solve these? In the above case, you can do a simple trial and error
solution to find that x1 = 2, x2 = 1 and x3 = 3, although this would be virtually
impossible if there were hundreds of equations. Remember, there is one equation for
every grid block in a linearised pressure equation in reservoir simulation (although
there are lots of zeros in the A matrix). This implies that we need a clear
numerical algorithm that a computer can work through and solve the linear
equations. The overview of approaches to solving these equations is discussed
in the next section.

5.2 General Methods for Solving Linear Equations

Firstly, let us note that there is a vast literature on solving sets of linear equations
and many books on the underlying theory and on the numerical techniques have
appeared. Indeed, petroleum reservoir simulation has led to the development of
some of these numerical techniques. We will not cover much of this huge subject
but we will cover enough that the student can appreciate - rather than understand in
detail - how the linear equation are solved in a simulator.

34

Numerical Methods in Reservoir Simulation

Basically, there are two main approaches for solving sets of linear equations involving
direct methods and iterative methods as follows:
(i) Direct Methods: this group of methods involves following a specific algorithm
and taking a fixed number of steps to get to the answer (i.e. the numerical values
of x1, x2 .. xn). Usually, this involves a set of forward elimination steps in order to
get the equations into a particularly suitable form for solution (see below), followed
by a back substitution set of steps which give us the answer. An example of such
a direct method is Gaussian Elimination.
(ii) Iterative Methods: in this type of method, we usually start off with a first guess
(or estimate) to the solution vector, say x(o) . Where we take this as iteration zero,
v = 0. We then have a procedure - an algorithm - for successively improving this
guess by iteration to obtain, x(1) x(2) x(3) .... x() etc. If it is successful, then
this iterative method should converge to the correct x as increases. The solution
should get closer and closer to the correct answer but, in many cases, we cannot
say exactly how quickly it will get there. Therefore, iterative methods do not
have a fixed number of steps in them as do direct methods. Examples of iterative
methods are the Jacobi iteration, the LSOR (Line Successive Over Relaxation)
method, etc.
The following two sections will discuss each of these approaches for solving
sets of linear equations in turn.

5.3 Direct Methods for Solving Linear Equations

In fact, we will not present the details of any direct method for solving linear equations.
Instead, we will discuss a general outline of how these methods work. Starting with
the basic linear equation in matrix form:

A.x = b

(60)

where A in an n x n square matrix and x and b are n x 1 column vectors. We can


write the n x (n+1) augmented matrix of A and b coefficients as follows:

a11

a21

a31

.....
.....

.....

an1

a12

a13

a14

a22

a23

a24

a32

a33

a34

an 2

an 3

an 4

a15 ... a1n b1

a25 ... a2 n b2

a35 ... a3n b3

... ...
... ...

... ...

an 5 ... ann bn

(61)

Remember that any mathematical operation we perform on one of the linear equations
(e.g. multiplying through by x2) must be performed on both sides of the equation i.e.
Institute of Petroleum Engineering, Heriot-Watt University

35

on the aij and the bi. Therefore, suppose we can transform the above augmented matrix
into the following form (we do not say how this is done, just that it can be done):

c11

.....
.....

.....
0

c12

c13

c14

c22

c23

c24

c33

c34

c15 ... c1n e1

c25 ... c2 n e2

c35 ... c3n e3

... ...
... ...

... ...
0 cnn en
...

(62)

The C-matrix above is in upper triangular form i.e. only the diagonals and above
are non-zero (cij 0, if j i) and all elements below the diagonal are zero (cij = 0,
if i > j), as shown above. Now consider why this particular form is of interest in
solving the original equations. The reason is that, if we have formed the augmented
matrix correctly (i.e. doing the same operations to the A and b coefficients), then the
following matrix equation is equivalent to the original matrix equation:
i.e.

A.x = b

<=>

C.x = e

c12

c13

c14

c22

c23

c24

c33

c34

(63)

Therefore,

c11

.....
.....

.....
0

c15 ... c1n x1 e1




c25 ... c2 n x2 e2


c35 ... c3n x3 = e3

.. . .
.. . .

.. . .
0 cnn xn en
...

(64)

This matrix equation is very easy to solve, as can be seen by writing it out
in full as follows:

36

Numerical Methods in Reservoir Simulation

c11 . x1 + c12 . x2 + c13 . x3 + c14 . x4 + c15 . x5 ... + c1n . xn = e1


c22 . x2 + c23 . x3 + c24 . x4 + c25 . x5 ... + c2 n . xn = e2
c33 . x3 + c34 . x4 + c35 . x5 ... + c3n . xn = e3
c44 . x4 + c45 . x5 ... + c4 n . xn = e4
......
......

...
...

cn 1, n 1 . xn 1 + cn 1, n . xn = en 1
cnn . xn = en
(65)
We can easily solve this equation by back-substitution, starting from the equation n
which only has one term to find xn, this xn is then used in the (n-1) equation (which
has 2 terms) to find xn-1 etc. as follows:

cnn . xn = en

cn 1, n 1 . xn 1 + cn 1, n . xn = en 1

=>

=>

xn = en / cnn now use xn in ..


e c .x
xn 1 = n 1 n 1, n n use xn 1 , xn in ..
cn 1, n 1

cn 2, n 2 . xn 2 + cn 2, n 1 . xn 1 + cn 2, n . xn = en 2

e
- cn 2, n 1 . xn 1 cn 2, n . xn
=> xn 2 = n 2
etc.
cn 2, n 2

(66)
Hence, by working back through these equations in upper triangular form, we can
calculate xn, xn-1, , xn-2.... back to x1.

Institute of Petroleum Engineering, Heriot-Watt University

37

EXERCISE 4.

Solve
triangular matrix:
+ 1example
= upper
4 x -the1following
x 2 x simple
x + 1of
x an
5
1

4 x1 - 1x2
2 x 2 2 x3 +
2 x2
3 x3 +

2 x3
1x4
2 x3
4 x4
3 x3
2 x4

+ 1 x 4 + 1x 5 = 5
+ 2 x5 = 12
+ 1x4 + 2 x5 = 12
+ 1x5 = 30
+ 4 x4 + 1x5 = 30
1x 5 = 3
2 x 4 1x 5 = 3
3 x5 = 15
3 x5 = 15

Answer
Answer:

Answer
x1 = ..........; x2 = ..........; x3 = ..........; x4 = ..........; x5 = ..........
x1 = ..........; x2 = ..........; x3 = ..........; x4 = ..........; x5 = ..........

5.4 Iterative Methods for Solving Linear Equations

As noted above, the idea in an iterative method for solving a set of linear equations is
to make a first guess at the solution and then to refine it in a stepwise manner using a
suitable algorithm. This procedure should gradually converge to the correct answer.
We illustrate the idea of such a method using a very simple iterative scheme. Note
that this simple point iterative scheme will work for some of the examples we
try here but it is not one that is recommended for use in reservoir simulation.
However, it adequately illustrates the main ideas which you need to know for the
purposes of this part of the course.
Our simple scheme starts with the longhand version of a set of linear equations and,
for our purposes, we will just take a set of four linear equations as follows:

a11 . x1 + a12 . x2 + a13 . x3 + a14 . x4 = b1


a21 . x1 + a22 . x2 + a23 . x3 + a24 . x4 = b2
a31 . x1 + a32 . x2 + a33 . x3 + a34 . x4 = b3
a41 . x1 + a42 . x2 + a43 . x3 + a44 . x4 = b4
We can rearrange the above set of equations as follows:

38

(67)

Numerical Methods in Reservoir Simulation

x1 =

1
b1 ( a12 x2 + a13 x3 + a14 x4 )
a11

x2 =

1
b2 ( a21 x1 + a23 x3 + a24 x4 )
a22

x3 =

1
b3 ( a31 x1 + a32 x2 + a34 x4 )
a33

x4 =

1
b4 ( a41 x1 + a42 x2 + a43 x3 )
a44

(68)

The resulting equations are precisely equivalent to the original set although, if
anything, they look a bit more complicated. Why would we deliberately complicate
the situation? In fact, the above reorganised equations forms the basis for an
iterative scheme, which we can use to solve the equations numerically. Firstly, we
need to introduce some notation as follows:
Notation:
( )
xi(v)

xi

the solution for xi at iteration ; where is the iteration


counter.

x ( 0 ) , x ( )

the first guess and th iteration of the solution vector, x.

Err.

an estimate of the error in the iteration scheme from the


to the (+1) iteration

Tol

some small quantity which determines the acceptable


error in the iteration scheme; if Err. < Tol, then the
scheme has converged.

Using the above notation in equations 68 above, gives the following simple iteration
scheme:

Institute of Petroleum Engineering, Heriot-Watt University

39

x1( +1) =

1
b1 ( a12 . x2( ) + a13 . x3( ) + a14 . x4( ) )
a11

x2( +1) =

1
b2 ( a21 . x2( ) + a23 . x3( ) + a24 . x4( ) )
a22

x3( +1) =

1
b3 ( a31 . x1( ) + a32 . x2( ) + a34 . x4( ) )
a33

x4( +1) =

1
b3 ( a41 . x1( ) + a42 . x2( ) + a43 . x3( ) )
a44

(69)

This iteration scheme may now be applied as follows:


(i)

Make an initial guess at the solution, iteration


= 0: x ( 0 ) = x1( 0 ) , x2( 0 ) , x3( 0 ) , x4( 0 )

(ii)

Update the solution to the next iteration +1 using equations 69


4

Err. = xik +1 x ki

(iii) Estimate an Error term (Err.) by comparing the latest with the previous
i =1
iteration of the unknowns, e.g.:
4

Err. = xi +1 xi
i =1

(iv) Is Err. < Tol.


If yes - the scheme has converged; if no - go back to step (ii) and continue
the iterations.
This scheme is now illustrated with a practical example.
Example: Solve the following set of equations using the iterative scheme above.

3.1x1
0.2 x1

- 0.32 x2 + 0.5 x3
+ 2.1x2

= 13.92

+ 0.33 x3 + 0.21x4 = 5.63

0.23 x1 - 0.32 x2 + 4.0 x3


0.42 x1 + 0.22 x2 +

40

+ 0.1x4

0. 5 x 3

+ 0.3 x4
+ 5.2 x4

= 15.19
= 16.76

Numerical Methods in Reservoir Simulation

(0)
(0)
(0)
(0)
Take the first guess: x1 = 2, x2 = 2, x3 = 2, x4 = 2

Hint: reorganise the above equations as follows:

x1( +1) = (1 / 3.1) * 13.92 - (- 0.32 x2( ) + 0.5 x3( )

(0.20 x

(0.23x

( 0.42 x

x2( +1) = (1 / 2.1) * 5.63 x3( +1) = (1 / 4.0) * 15.19 x4( +1) = (1 / 5.2) * 16.76 -

( + 1)
1

+ 0.1x4( ) )

+ 0.33 x3( ) + 0.21x4( )

( +1)
1

( +1)
1

0.32 x2( +1)

))

+ 0.30 x4( )

+ 0.22 x2( +1) +

0.5 x3( +1)

))

))

EXERCISE 5.

Fill in the table below for 10 iterations using a calculator or a spreadsheet:


POINT ITERATIVE SOLUTION OF LINEAR EQUATIONS
Iteration
counter

First guess = 0
1
2
3
4
5
6
7
8
9
10

x at iter.
x1
x2
2.0
2.0

x3
2.0

x4
2.0

Iteration
counter x at iter.

x1
x2
x3
x4
In some First
cases,
it may
guess
= 0be possible
2.0 to develop
2.0 improved
2.0 iteration
2.0 schemes by using
the latest information 1that is available.
For
example,
in
the
above
iteration scheme,
4.309677 1.97619
3.6925
2.784615
(+1)
2
4.008926
1.411795
3.498943
2.436331
we could use the very latest value of x1
when we are calculating x2(+1) since we
3
3.993119
1.505683
3.497206
already have this quantity. Likewise, we can use both
x1(+1) 2.503112
and x2(+1) when we
4
4.000937 1.500783 3.500617 2.500584
(+1)
calculate x3
etc as shown below:
5
3.999963 1.499755 3.499965 2.499832
6
3.999986 1.500026 3.499995 2.500017
7
4.000003 1.5
3.500002 2.500001
8
4
1.499999 3.5
2.5
Converged 9
4
1.5
3.5
2.5
10
4
1.5
3.5
2.5
Note - converges fully at k = 9 iterations.

Institute of Petroleum Engineering, Heriot-Watt University

41

x1( +1) = (1 / 3.1) * 13.92 - (- 0.32 x2( ) + 0.5 x3( )

(0.20 x

(0.23x

( 0.42 x

x2( +1) = (1 / 2.1) * 5.63 x3( +1) = (1 / 4.0) * 15.19 x4( +1) = (1 / 5.2) * 16.76 -

( + 1)
1

( +1)
1

( +1)
1

+ 0.1x4( ) )

+ 0.33 x3( ) + 0.21x4( )

0.32 x2( +1)

))

+ 0.30 x4( )

+ 0.22 x2( +1) +

0.5 x3( +1)

))

))

x1( +1) , x2( +1) , x3( +1) => underlined terms, imply they are at the latest time

available).

When this is done, we find the following results:


IMPROVED POINT ITERATIVE SOLUTION OF LINEAR EQUATIONS (Using
latest information)
- see spreadsheet CHAP6Ex5.xls - Sheet 2
Iteration
counter

First guess = 0
1
2
3
4
5
Converged 6
7
8
9
10

x at iter.
x1
2
4.309677
4.021247
3.999376
3.999973
3.999998
4
4
4
4
4

x2
2
1.756221
1.495632
1.500005
1.499974
1.5
1.5
1.5
1.5
1.5
1.5

x3
2
3.540191
3.501408
3.500161
3.499997
3.5
3.5
3.5
3.5
3.5
3.5

x4
2
2.460283
2.498333
2.500035
2.500004
2.5
2.5
2.5
2.5
2.5
2.5

Note - converges fully at = 6 iterations.

Clearly, comparing this table with the previous one using the simple point iterative
method, we see that this method does indeed converge quicker. Although this
is just one example, it is generally true that using the latest information in an
iterative scheme improves convergence.
Notes on iterative schemes: there are several point to note about iterative
schemes, which we have not really demonstrated or explained here. However,
you can confirm some of these points by using the supplied spreadsheets.
These points are:
(i) Iterative solution schemes for linear equations are often relatively simply to apply
- and to program on a computer (usually in FORTRAN). If you study the supplied
spreadsheets (CHAP6Ex5.xls), you can see that they are quite simple in structure
for this set of equations.
42

Numerical Methods in Reservoir Simulation

(ii) The convergence rate of an iterative scheme may depend on how good the initial
guess (x(0)) is. If you want to demonstrate this for yourself, run the spreadsheet
(CHAP6Ex5.xls) with a more remote initial guess. Here is the same example as
that presented above with a completely absurd initial guess:
POINT ITERATIVE SOLUTION OF LINEAR EQUATIONS - Bad First Guess
(you can use CHAP6Ex5.xls to confirm these results)
Iteration
counter

First guess = 0
1
2
3
4
5
6
7
8
9
10
11
12
Converged 13
14
15

x at iter.
x1
900
-1017.45
63.79526
55.59796
1.890636
4.326953
4.090824
3.987986
4.001064
4.000117
3.999963
4.000004
4
4
4
4

x2
-2000
-2454.69
406.2002
-20.1756
-2.67691
2.668945
1.389409
1.49622
1.502872
1.499593
1.500013
1.500006
1.499999
1.5
1.5
1.5

x3
456
-1932.95
-131.922
4.491703
-0.53117
3.538073
3.519613
3.491895
3.500729
3.500025
3.499982
3.500003
3.5
3.5
3.5
3.5

x4
23000
-28.7
375.1144
-6.43019
0.84584
3.234699
2.420476
2.495457
2.50191
2.499722
2.500005
2.500004
2.499999
2.5
2.5
2.5

(iii) We cannot tell in advance how many iterations may be required in order to
converge a given iterative scheme. Clearly, from the results above, a good initial
guess helps. In reservoir simulation, when we solve the pressure equation, a good
enough guess of the new pressure is the values of the old pressures at the last
time step. If nothing radical has changed in the reservoir, then this may be
fine. However, if new wells have started up in the model or existing wells have
changed rate very significantly, then the pressure at last time step guess may
not be very good. However, with a very robust numerical method, convergence
should still be achieved.
(iv) It helps in an iterative method to use the latest information available. This was
demonstrated in the iterative schemes in spreadsheet CHAP6Ex5.xls.

5.5 A Comparison of Iterative and Direct Methods for Solving Linear


Equations

In this Chapter, we have described two general methods - direct and iterative - for
solving the linear equations which arise when we discretise the flow equations of
reservoir simulation. We have not indicated which of these methods is usually best
for reservoir simulation problems. This is because, it depends on the problem. The
final numerical problem which is solved in a reservoir simulator could be quite small
and simple, or it could be very large and have some intrinsically difficult numerical
problems within it.

Institute of Petroleum Engineering, Heriot-Watt University

43

In any numerical computational method, such as those for solving a set of linear
equations, we need to have some way of calculating the amount of work involved.
To some extent this depends on the computer architecture since new generation parallel
processing machines are becoming available as discussed in Chapter 1. However, we
will take quite a simple view based on a conventional serial processing machine and
will define computational work as simply the number of multiplications and divisions
(addition and subtraction is cheap!). In this way, we can define the amount of work
required for any given direct and iterative scheme for solving linear equations. We
present results without any proof for two such schemes called the BAND (a direct
method) and LSOR (Line Successive Over-Relaxation; an iterative method) schemes.
For a 2D problem with NX and NY grid blocks in the x and y directions, respectively
(where we choose NY < NX)
BAND (direct),

Work, WB

NX.NY3

(70)

LSOR (iterative),

Work, WLSOR

NX.NY. Niter

(71)

where Niter is the number of iterations until convergence is reached. It is already


quite clear that the amount of work for the direct method - which only has to be done
once - is larger than that for a single iteration of the iterative method. However, we
do not know in advance the number of iterations, Niter. We illustrate some typical
results for these two methods with a simple exercise below.

EXERCISE 6.

Which is best method (i.e. that requiring the lowest amount of computational work),
BAND or LSOR, for the following problems?
(i) NX = 5,

NY = 3,

(ii) NX = 20, NY = 5,

Niter = 50 (a small problem)


Niter = 50

(iii) NX = 100, NY = 20, Niter = 70


(iv) NX = 400, NY = 100, Niter = 150

44

NX

NY

Niter

5
20
100
400

3
5
20
100

50
50
70
150

BAND
WB

LSOR,
WLSOR

Comment

Numerical Methods in Reservoir Simulation

We can summarise our comparison between direct and iterative methods for solving
the sets of linear equations that arise in reservoir simulation as follows:
(i) A direct method for solving a set of linear equations has an algorithm that
involves a fixed number of steps for a given size of problem. Given that the
equations are properly behaved (i.e. the problem has a stable solution), the direct
method is guaranteed to get to the solution in this fixed number of steps. No first
guess is required for a direct method.
(ii) An iterative method on the other hand, starts from a first guess at the solution (x(o))
and then applied a (usually simpler) algorithm to get better and better approximations
to the true solution of the linear equations. If successful, the method will converge
in a certain number of iterations, Niter, which we hope will be as small as possible.
However, we cannot usually tell what this number will be in advance. Also, in some
cases the iterative method may not converge for certain types of difficult problem.
We may need to have good first guess to make our iterative method fast. Also, it
often helps to use the latest computed information that is available (see example
above).
(iii) Usually the amount of work required for a direct method is smaller for
smaller problems but iterative methods usually win out for larger problems. For
an iterative method, the amount of work per iteration is usually relatively small
but the number of iterations (Niter) required to reach convergence may be large and
is usually unknown in advance.

Institute of Petroleum Engineering, Heriot-Watt University

45

EXERCISE 7.

The linear equations which arise in reservoir simulation may be solved by a direct
solution method or an iterative solution method. Fill in the table below:

Direct solution method

Iterative solution method

1.

1.

2.

2.

1.

1.

2.

2.

Give a very
brief description
of each method

Main advantages

Main
disadvantages

46

Numerical Methods in Reservoir Simulation

6. DIRECT SOLUTION OF THE NON-LINEAR EQUATIONS OF


MULTI-PHASE FLOW
6.1 Introduction to Sets of Non-linear Equations

We noted in Section 4 above that the equations which we obtain when we discretise
the two-phase flow equations are actually non-linear in nature. However, the strategy
we discussed above (the IMPES method) involved tackling the problem almost as
if it were a linear set of equations since we time-lagged the coefficients to reduce
the problem to a linear set of equations. Then we repeated this process - we applied
repeated iterations - until it (hopefully) converged. Therefore, we took a non-linear
problem and solved it as if it were a series of linear problems.
In this section, we will introduce the general idea of solving sets of non-linear equations
numerically and indicate how this can be applied to the two-phase flow equations.
We will do this in a simplified manner that shows the basic principles without going
into too much detail. Firstly, compare the difference between solving the following
two sets of two equations, a set of 2 linear equations:

2 x1 + x2 = 10
3 x1 x2 = 5

(72)

and a set of 2 non-linear equations:

x1

2 x2 .e x1 = 2

x2 .( x1 )2 x1 .( x2 ) = 5
2

(73)

It is immediately obvious that the second set of non-linear equations is more difficult.
The first set of linear equations can be rearranged easily to show that x1 = 3 and x2 =
4. However, it is not as straightforward to do the same thing for the non-linear set
of equations. As a first attempt to solve these, we might try to develop an iterative
scheme by rearranging the equations as follows
( )

x1( +1) = 2 2 x2( ) .e x1


x

( +1)
2

x2( ) .( x1( ) )2 +5
x1( )

(74)

where, as before, denotes the iteration counter.


The actual solution in this case is, x1 =2.51068, x2 = 3.144089. Applying the above
iterative scheme with a first guess, x1(0) = x2(0) = 1.0, gives the results shown in Table
2 below (where we have removed some of the intermediate iterations).
Table 2: Non-linear scheme applied to solution of equations 73 using the point iterative
scheme of equation 74 (spreadsheet, CHAP6Ex5.xls - Sheet 3)

Institute of Petroleum Engineering, Heriot-Watt University

47

Iteration
counter

First guess = 0
1
2
3
4
....
9
10
....
25
26

x at iter.
x1
1
2.735759
2.317673
2.575338
2.454885
....
2.513337
2.508958
....
2.510682
2.510681

x2
1
2.44949
2.920421
2.987627
3.104133
....
3.141814
3.144173
....
3.144089
3.144089

You can check the results in Table 2 or experiment with other first guesses using
Sheet 3 of spreadsheet CHAP6Ex5.xls. We note that the convergence rate in Table
2 is not very fast but, in this case, it does reach a solution. In general, it is usually
very difficult to establish for certain whether a given scheme will converge for nonlinear equations although there is a large body of mathematics associated with the
solution of such systems (which is beyond the scope of this course).

6.2 Newtons Method for Solving Sets of Non-linear Equations

We take a step back to the solution of a single non-linear equation such as:

x 2 .e x = 0.30

(75)

which has the solution x = 0.829069 (you can verify this by calculating x 2 .e x
and making sure it is 0.30 - to an accuracy of ~ 4.6x10-8). This equation can be
represented as:

f ( x ) = 0 where f ( x ) = x 2 .e x 0.30

(76)

and the solution we require is the value of x for which the function f(x) is zero.
Expand f(x) as a Taylor series as follows:

f ( x ) f ( xo ) + x. f ' ( x0 ) +

x 2 ' '
. f ( x0 ) + ...
2

(77)

Thinking in terms of an iteration scheme (x() x(+1)) as a basis for calculating


better and better guesses of the solution of f(x) = 0, we can rewrite the Taylor series
above as:

f ( x ) f ( x ( ) ) + ( x ( +1) x ( ) ). f ' ( x ( ) )

(78)

where we have neglected all the second order and higher terms. Since we require
the solution for f(x) = 0, then we obtain:

f ( x ( ) ) + ( x ( +1) x ( ) ). f ' ( x ( ) ) = 0

48

(79)

Numerical Methods in Reservoir Simulation

which can be rearranged to the following algorithm to estimate our updated guess,
x(+1) as follows:

x ( +1) = x ( )

f ( x ( ) )
f ' ( x ( ) )

(80)

This equation is the basis of the Newton-Raphson algorithm for obtaining better and
better estimates of the solution of the equation, f(x) = 0. Note that we need both a first
guess, x(0), and also an expression for the derivative f ' (x()) at iteration . In the simple
example in equation 76, we can obtain the derivative analytically as follows:

df ( x )
= f ' ( x ) = 2 x.e x x 2 .e x
dx

(81)

Therefore the Newton-Raphson algorithm for solving equation 75 above is as


follows:

( + 1)

=x

( )

( x ( ) )2 .e x ( ) 0.30

- ( ) x ( )
( )
2
2 x .e
( x ( ) ) .e x

(82)

The point iterative method of the previous section and the Newton-Raphson method
of equation 82 have both been applied to the solution of equation 75 (see Sheet 4
of spreadsheet CHAP6Ex5.xls for details). The results are shown in Table 3 for a
first guess of x(0) = 1.
Table 3: Comparison of the point iterative and Newton-Raphson methods for solving
non-linear equation 75; (see spreadsheet CHAP6Ex5.xls - Sheet 4 for details)
Iteration
Number

Value of x at iteration

Point
Iteration
method
x()

NewtonRaphson
method
x()

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

1
0.903042
0.860307
0.84212
0.834497
0.831322
0.830003
0.829456
0.82923
0.829136
0.829097
0.82908
0.829074
0.829071
0.82907
0.829069
0.829069

1
0.815485
0.825272
0.82806
0.828805
0.829
0.829051
0.829064
0.829068
0.829069
0.829069
0.829069
0.829069
0.829069
0.829069
0.829069
0.829069

Institute of Petroleum Engineering, Heriot-Watt University

49

The correct solution to equation 75 is x = 0.829069. The results in Table 3 show


that the Newton-Raphson method converges to this solution (to an accuracy better
than 10-6) in 9 iterations, whereas it takes 15 iteration for the point iterative method
to converge at the same level accuracy. This performance can vary quite a bit from
problem to problem but the Newton-Raphson method is generally better if the derivative
term, f '(x()), is not too close to zero. When this derivative gets too small, it can be
seen that the second term in equation 80 would start to get very large or blow up,
as it is sometimes described. We will not go into details about the convergence
properties of the Newton-Raphson but it is meant to be quadratic meaning that the
error should decrease quite rapidly.
We now go on to see how the Newton-Raphson method can be applied to sets of
non-linear equations. Returning to the set of two non-linear equations in the previous
section, we note that another way of writing this set of equations in a general way
is as follows:

F1 ( x1 , x2 ) = 0
F2 ( x1 , x2 ) = 0

(83)

where, in the case of equation 73 above, these functions would be given by:

F1 ( x1 , x2 ) = x1

2 x2 .e x1 2

F2 ( x1 , x2 ) = x2 .( x1 )2 x1 .( x2 ) + 5
2

(84)

The problem is then to find the values of x1 and x2 that make F1 = F2 = 0. In this
section, we present Newtons method for the solution of sets of non-linear equations.
Basically, we will simply state the Newton-Raphson algorithm without proof (although
it will resemble the simple form of the Newton-Raphson method above) for sets of
non-linear equations. We will then illustrate what this means for the example of the
set of two non-linear equations above (equation 73).
Definitions:
As before,
( )
andx (x+(1)+1=) the solution vectors at iteration levels and (+1)
x (x) and

New terms
are:
( )

(
x) ) ) ( +1)
F(F
xx(( ) and
x

( )

x)
x (F
( x ( ) )

= the number of non-linear equations (and hence


unknowns, x1, x2, .... xN)
= the vector of function values, F1, F2 .. FN at x()

the"true"
"true"solution
solutionofofthe
thesetsetofofnon
non
- linearequations
equations
- linear
F(Fx(()x) =) =0 0forforthe
x
50

F( x ) = 0 for the "true" solution of the set of non - linear equations

Numerical Methods in Reservoir Simulation

That is:

F1 ( x1 , x2 , x3 ,....., x N )

F2 ( x1 , x2 , x3 ,....., x N )

F3 ( x1 , x2 , x3 ,....., x N )
( )
F( x ) =

.....

.....

F ( x , x , x ,....., x )
N
N 1 2 3

J ( x ( ) )

(85)

= the NxN Jacobian matrix defined as follows:

F1 ( x ( ) ) F1 ( x ( ) ) F1 ( x ( ) )
F1 ( x ( ) )
......

x2
x3
x N
x1

F2 ( x ( ) ) F2 ( x ( ) ) F2 ( x ( ) )
F2 ( x ( ) )
......

x2
x3
x N
x1

J ( x ( ) ) =
F3 ( x ( ) ) F3 ( x ( ) ) F3 ( x ( ) )
F3 ( x ( ) )
......

x
x N
1
2
3

.....

.....

( )
( )
( )
( )
FN ( x ) FN ( x ) FN ( x ) ...... FN ( x )
x1
x2
x3
x N (86)

[ J ( x )]
( )

= the inverse of the Jacobian matrix. (Recall that the inverse of a


matrix A is denoted by A-1 and by definition, A-1 .A = I, where I is the identity matrix
- all diagonals 1 and all other elements zero - see below). The above matrix is also
an NxN matrix with the property:

[ J ( x )] .[ J ( x )] = I, the identity matrix


[ J ( x )] .[ J ( x )] = I, the identity matrix
( )

( )

( )

( )

1 0 0 0 0
10 0 1 0 0 0 0 0 0
0 1 0 0 0
0 0 1 0 0
00 0 0 1 0 0 1 0 0
1
That is J ( x ( ) ) . J ( x ( ) ) = I =
0 0 0 1 0
1
0 0 0 0 1
( )
( )
J(x ) . J(x ) = I =
0 .......
0 0 0 1

.......
.......

.......
0 0 0 0 0

0 0 0 0 0

] [
] [

Institute of Petroleum Engineering, Heriot-Watt University

...... 0
......
...... 0 0

......
...... 0 0
......
...... 0 0
(87)

......
...... 0 0
...... 0

...... 1

...... 1

51

Using all of the above definitions, the Newton-Raphson algorithm for a set of nonlinear equations, F( x ) = 0, is given by the following expression:

x (+1) = x () J ( x () ) .F( x () )
1

(88)

We first demonstrate how this is applied to a simple example (equations 73) before
going on to show how it is applied to the more complicated non-linear sets of
equations which arise in the fully implicit discretisation of the reservoir simulation
equations.
Example: Returning to the simple example, where N = 2 (rearranged equations
73):

F1 ( x1 , x2 ) = x1

2 x2 .e x1 2

F2 ( x1 , x2 ) = x2 .( x1 )2 x1 .( x2 ) + 5
2

(89)

In the notation developed above, these equations become:

x1 + 2 x2 .e x1 2
F1 ( x1 , x2 )
F( x ) =

=
2
x2 .( x1 )2 x1 .( x2 ) + 5
F2 ( x1 , x2 )

(90)

We note that the 2x2 Jacobian matrix is given by:

F1 ( x ( ) )

x1
( )
J(x ) =

F2 ( x ( ) )

x1

F1 ( x ( ) )

x2

F2 ( x ( ) )

x2

(91)

where each of the elements can be evaluated analytically to obtain:

1 2 x ( ) .e x1( )
2

J ( x ( ) ) =

2 x ( ) . x ( ) x ( ) 2
(2 )
2 1

(2.e )
x1( )

) (( x

) 2x

( ) 2
1

( )
1

. x2( )

Hence, the Newton-Raphson method for this simple system becomes:

52

(92)

Numerical Methods in Reservoir Simulation

x ( + 1) = x ( )

1 2 x ( ) .e x1( )
2

2 x ( ) . x ( ) x ( ) 2
(2 )
2 1

(2.e )
x1( )

) (( x

) 2x

( ) 2
1

( )
1

. x2( )

x ( ) + 2 x ( ) .e x1( ) 2
2
1

2
x2( ) .( x1( ) )2 x1( ) .( x2( ) ) + 5

(93)

[ ( )]
( )

which we could solve if we knew how to invert the Jacobian matrix to find J x
This is a detail which we dont need to know for this course but it can be done.
Indeed, for a 2x2 matrix - such as in the above case - it is quite easy to work out the
inverse. The inverse of a simple 2 x 2 matrix A given by:
A = a b

c d

is well known to be
A-1 =

1
d b

( ad bc) c a

This can easily be proven by multiplying out there matrices and showing that
A-1 .A = I. The factor (ad-bc) is known as the determinant of the matrix and it must be
non-zero for A-1 to exist. Rather than using equation above to write out the analytical
form of the J ( x ( ) ) matrix we first simply numerically evaluate the matrix J at a
1
given iteration and apply equation above no get J . This is done in the spreadsheet
Chap6Exxx.xls where results are shown.

[ J ( x )]
(k )

= ***

[]

(94)

which allows us to apply the Newton-Raphson method directly to our simple example.
For details see Sheet 5 spreadsheet ***.xls. The results are shown in Table 6.xx.
(NOTES & SPEADSHEET STILL UNDER CONSTRUCTION)

Institute of Petroleum Engineering, Heriot-Watt University

53

6.3 Newtons Method Applied to the Non-linear Equations of TwoPhase Flow

We now return to the discretised equations of two phase flow which we noted at the
time, formed a set of non-linear equations. The simplified form of these equations
is as follows (see Section 4.1):
Pressure (equation 50):

( (S ))
T

n +1
o

i 1 / 2

n +1
i 1

(Son +1 )
T

x 2

i 1 / 2

( (S ))
+
n +1
o

i +1 / 2

(S n + 1 )
P n +1 + T o
i
x 2

i +1 / 2

Pin++11 = 0

Saturation (Form B - equation 54):


Soin+1 Soin

n +1

t o ( So )


x 2

( Son +1 )
n +1
n +1

P
P
( i +1 i ) o x 2

i +1 / 2

i 1 / 2

Given that the unknowns that we are trying to find are Pi


then we can write the above equations in general form as:

n +1
n +1

P
P
(i
i 1 ) = 0

n +1

, Sin +1 i = 1,2, NX. ,

FP, i ( S n +1 , P n +1 ) = 0

FS , i ( S n +1 , P n +1 ) = 0

(95)

FP, i ( S n +1 , P n +1 ) andn +F1S , i (nS+n1+1 , P n +1 ) n +1 n +1


where the two non-linear equations, FP, i ( S , P ) and FS , i ( S , P ) , arise from
the pressure (P) and saturation (S) equations, respectively, as given above. The
vectors of unknowns, S n +1 and P n +1, nare
+1 given nby
+1 :

and P

n +1
1

S n +1

54

S
n +1
S2
n +1
S3
...

...
S n +1
= i 1
Sin +1
n +1
Si +1

...
...

...
S n +1
NX

and

P n +1

P1n +1
n +1
P2
n +1
P3
...

...
P n +1
= i 1
Pi n +1
n +1
Pi +1

...
...

...
P n +1
NX

(96)

Numerical Methods in Reservoir Simulation

FP, i ( S n +1 , P n +1 ) and FS , i ( S n +1 , P n +1 )

n +1

n +1

n +1

n +1

and FS , i S , P
However, the given FP, i S , P
at grid block i only depend
on the quantities Sin+11 , Sin +1 , Sin++11 and Pi n1+1 , Pi n +1 , Pi n+1+1 and, rather than on all of the
other saturations and pressures in the system, since these are the nearest neighbours
+1
n +1
n +1
Sin+11equations
, Sin +1 , Sinabove.
, Pi n+1+1
coupled together in the
+1 and Pi 1 , Pi
We may also write one total unknows vector Xn+1 using a combination of the saturation
and pressure vectors as follows:

X n +1

S1 n +1
n +1
P1
S n +1
2
P2 n +1

=
S n +1
i
Pi n +1

S n +1
NX
n +1
PNX

and the equation to be solved is then F(Xn+1)=0


The saturations and pressures are coupled together to their nearest neighbours through
the discretisation equations (50 and 54 above). The general form of the solution using
the Newton iteration is then to take a starting guess (iteration, =0) Xn+1(0) and then
apply the formulation as above to obtain:

X n +1 ( +1) = X n +1( ) + J ( X )( )

. F ( X ( ) )

It is rather more complex to constuct the Jacobian matrix J ( X ) of derivatives but


1
it can be done. We also need to invert this matrice to obtain J in order to apply
the above algorithm. However, in practice, there are various methods that try to
1
construct the J matrix more directly, often in an approximate manner. Note that
the Jacobian matrix is very sparse since there are just nearest neighbour interaction
(as was the case for the matrices associated with single phase flow discussed earlier
in this Chapter). A consequence of this is that the inverse matrix is also quite sparse
and there are many numerical techniques available to solve such problems.

[]

[]

In this course, we will not give any more detail on the solution of the non-linear
equations which arise in reservoir simulation.

Institute of Petroleum Engineering, Heriot-Watt University

55

7 NUMERICAL DISPERSION - A MATHEMATICAL APPROACH


7.1 Introduction to the Problem

In Chapter 4, we discussed the physical idea of numerical dispersion, which is also


sometimes referred to as numerical diffusion. Numerical dispersion is essentially
the artificial spreading or diffusion of a front - e.g. a waterfront in a water/oil
displacement - due to the coarse grid used in the simulation. It is a numerical effect
and it can be reduced by taking a finer grid. We note that in real reservoirs there are
some real dispersive physical mechanisms which result in the spreading of fronts.
These arise due to the effects of capillary pressure and also from the interaction of the
fluid flow with small scale (cm - m) permeability heterogeneity of the reservoir rock.
Indeed, there are also some quite complex interactions between the capillary forces
and the small scale heterogeneity that also lead to types of physical spreading in the
reservoir. In an ideal calculation, we would take a sufficiently fine grid that the level
of physical (i.e. real) spreading or diffusion was correctly represented by our grid;
i.e. the numerical diffusion would be less that the physical diffusion. This is almost
never possible in a field scale simulation, although it can be achieved in modelling
laboratory scale experiments on flows through small rock samples or bead packs.
For such a fine scale simulation, the level of numerical diffusion can realistically be
made much less than that from physical sources.
In this section, we return to the issue of numerical dispersion - a term which we will
now use interchangeably with numerical diffusion. Indeed, diffusion is often
referred to as quite a specific physical effect which is described by certain well-known
equations which are, not surprisingly, known as diffusion equations. For example,
the simplified pressure equation derived in Chapter 5, equation 27, is an example of
a diffusion equation. This equation has the form:

2 P
P
= Dh 2
t
x

(97)

where Dh is the hydraulic diffusivity (Dh = k/( cf ..)) and this is a standard form
of the classical diffusion equation. We now consider how the effects of a grid can
lead to diffusion-like terms when we try to solve the flow equations numerically.

2 P

In the above equations, it is specifically the 2 term that is the dispersive or


x
diffusive part.

7.2 Mathematical Derivation of Numerical Dispersion

In order to show mathematically how a diffusive term arises when we solve certain
transport equations numerically, we use a simple form of the two-phase saturation
equation. In 1D, the transport equation for a simple waterfront (with Pc = 0 and zero
gravity) is the well-known fractional flow equation (see Chapter 2) as follows:

S
f
w = v w
t
x

(98)

where Sw is the water saturation and fw ( Sw ) is the fractional flow of water

56

qw
fw ( Sw ) =

qw + qo

Numerical Methods in Reservoir Simulation

fw ( Sw )

qw
) which is a function only of water saturation, Sw. We can
qw + qo

( fw ( Sw ) =

differentiate fw ( Sw ) by the chain rule to obtain:

fw ( Sw )

f S
S
f
w = v w = v w w
ft S = xqw
Sw x

w( w)
qw + qo

(99)

f
v w
where the termfw ( Sw)Sw is a non-linear term giving the water velocity, v w (Sw ) ,
which is also a function of water saturation: that is:

f
v w (S w ) = v w
Sw

fw (Sw )

(100)
To simplify the problem even further for our purposes here, we take a straight line

fw

fractional flow, fw (Sw ) , which means that Sw is a constant and, hence, so is


the water velocity, Vw. Hence, the simple 1D transport equation for a convected
waterfront is:

fw

S
w S Sw

= wv w

t
x

(101)

where, as noted above, vw is now constant. This equation describes the physical
situation illustrated schematically in Figure 14.
t1

Governing equation
Sw
Sw
= - Vw
t
x

t2

Figure 14
The advance of a sharp
saturation front governed
by the transport equation
with a constant velocity vw.

Sw
Water
0

Vw
x1

Oil
x2

Vw = constant
Vw =

x2 - x1
t2 - t1

Starting from the transport equation 101 above, we can easily apply finite differences
using our familiar notation (Chapter 5) to obtain:
n
Sn +1 Swi

Swi Swi 1
wi
= vw
+ Error terms

t
x

(102)

where we have used the backward difference (sometimes referred to as the upstream

Sw

difference) to discretise the spatial term, x . Note that we have not yet specified
the time level of the spatial terms. If we take these at the n time level (known), then it

Institute of Petroleum Engineering, Heriot-Watt University

57

would be an explicit scheme and if we take them at the n+1 time level (unknown), it
would be an implicit scheme. It does not matter for our current purposes here, since
we are principally interested in the Error terms which are indicated in equation
102. To determine what these terms look like, we need to go back to the original
finite differences based on Taylor expansion of the underlying function, Sw(x,t), in
this case.
We can expand Sw (x,t) either in space (x) or in time (t) as follows:
SPACE

2
2
S x Sw
Sw (x, t ) = Sw (x 0 ) + x. w +

+ higher order terms


x
2 x 2
(t fixed)

(103)

t 2 2S

w
w
+
TIME Sw ( x, t ) = Sw ( t 0 ) + t.
2 + higher order terms
t

t
(x fixed)
(104)

We can rearrange each of the above equations 103 and 104 to obtain the finite

S
S
and

difference approximations for the x


t terms with the leading error
terms as follows (now ignoring the higher order terms):

S(x, t ) S(x 0 ) x 2S
S
SPACE
2
x

TIME

2 x

2
S S(x, t ) S( t 0 ) t S

2 t 2
t
t

(105)

(106)

To return to our usual notation, we make the identities:

S( x, t ) Sin +1
S( x 0 ) Si 1
S( t 0 )

Sin

(107)

Equation 105 now becomes:


2
S Si Si 1 x S

x x 2 x 2

(108)

and equation 106 becomes:


n +1
n
2
S Si Si t S


t t 2 t 2

58

(109)

Numerical Methods in Reservoir Simulation

We now substitute the above finite difference approximations into the governing
equation 102 with their leading error terms as follows:

Sn +1 Sin t 2S
Si Si 1 v w .x 2S
i

w
t 2 t 2
x
2 x 2 (110)
Collecting the error terms together on the RHS gives:

Sn +1 Sin
Si Si 1 v w .x 2S t 2S
i
v

w
t
x
2 x 2 2 t 2
Error term

(111)

The error term in the finite difference scheme is now clear and is shown in equation
111. However, it still needs some simplifying since it is a strange mixed term with

2S
2S
2S
2 and 2
both x
t terms in it. We now want to eliminate the t 2 term
and this is done by using the original governing equation.

2S
Sw
S

= vw w
2 , first note from the original equation 101 that
x
t
t
Sw
with respect to t as follows:
and we can then differentiate
t
To obtain

Sw 2Sw
S

= 2 = vw w
t t t
t x

(112)

Thus, we can rearrange the RHS of equation 112 as follows:

vw

Sw
S

= vw w
t x
x t

(113)

We now return again to the governing equation (equation 101) and use it for w
t
in equation 113 to obtain:

vw

2
Sw

Sw
2 Sw
=

=
v
v
v

w
w
x 2
x t
x
x

(114)

and hence:
2
2 Sw
2 Sw
=
v
2

w
t
x 2

(115)

We now substitute this expression into the error term in equation 111 above to obtain
the following:
Institute of Petroleum Engineering, Heriot-Watt University

59

v w .x 2S t 2S v w .x 2S v 2w t 2S

+
=

2 x 2 2 t 2
2 x 2
2 x 2
v .x v 2w t 2S
= w
+

2
2 x 2

(116)

The finite difference equation with its error term in equation 111 now becomes:

Sin +1 Sin
Si Si 1 v w .x v 2w t 2S
v
+

w
t
x 2
2 x 2

(117)

In this equation, we now see that the form of the error term is exactly like a diffusive
term i.e. it multiplies a (2Sw/x2) term. Hence we identify the level of numerical
dispersion or diffusion, Dnum, arising from our simple finite difference scheme as:

v .x v 2w t
x v w t
D num = w
+
+
= v w .

2
2
2
2
x

If D num = v w .
+
For this case: 2

D num
v w t <<
x

x v w t
D num = v w .
+

2
(118) 2

v w t
, we can take such a small time step that v w t << x
2

v w .x
2

(119)

We can now solve two problems as follows


(i) Firstly, we may apply an explicit finite difference method to obtain an accurate
solution of the following convection-dispersion equation.

2 Sw
Sw
Sw
=

v
+
D
.

2
w
t
x
x

(120)

That is, where the numerical dispersion is much less than the physical dispersion due
to the explicit D term. If this solution is converged (i.e. does not change on refining
x and t) then we can plot this at a suitable time, t1, as shown in Figure 15.

Governing equation:

Sw
= - Vw
t

t = t1

Sw

Vw = water velocity

60

Sw
+D
x

Sw
x2

Figure 15.
Dispersed frontal displacement with a known and converged level of dispersion.

Numerical Methods in Reservoir Simulation

n +1
n
n
n
Swi

Swi
Swi
Swi
1
(ii) Now take only the explicit finite difference approximation
tov wthe

=
convection

t
x
equation only, that is:
n
n
n
Sn +1 Swi

Swi
Swi
v .t n
1
n +1
n
n
wi
= vw
which gives Swi
= Swi
w (Swi
Swi
1 )

t
x
x.

Solve this with a relatively coarser grid, x, but with a fine time step thus predicting
x n
vwv.
n +1
n
n
w .t
=

Swi Swi
S
S

Dwinum to be
level of dispersion to deliberately match that in
wi
this
1)
2x..( Choose

part (i) above i.e. choose x such that Dnum = D. Plot out a saturation profile like
that in Figure 15 at the same time t1 and compare these.

CLOSING REMARKS

In this Chapter, we have introduced the student to finite difference approximations


of the partial differential equations (PDEs) that describe both single- and two-phase
flow through porous media. These discretised equations have led to systems of either
linear or non-linear equations which are then solved numerically. Methods for solving
these equations have been discussed in principle and some idea has been given of
how these are applied in practical reservoir simulation. At the end of the unit, some
discussion was presented on grid-to-grid flows and how these inter-gridblock properties
are averaged. A more mathematical of numerical dispersion was also given.

Institute of Petroleum Engineering, Heriot-Watt University

61

APPENDIX A: Some useful Matrix Theorems


This lists (without proof) some useful theorems from matrix algebra which underpin
much of the practical application work which we have described in this section.

A.x = b

The central problem which we are interested in is:

A.x = b
A.x = b where A is an NxN square matrix
1. If A.x = b actually has a solution, then A is said to be non-singular or invertible
Aand the following five conditions apply and are equivalent (i.e. if one of them hold,
they all hold):

I
A
(i) A 1 exists where A 1 . A = I ; I is the identity matrix which has 1s on the
diagonal and zeros elsewhere. I is like the number 1 in that I.A = A.I = A .
(ii) det( A ) 0 ; the determinant of the matrix is non-zero. This is a number found
A a=specific
A.I = Away.
by multiplying out the matrixI.in
(iii) There is no non-zero
A.x = 0vector x such that, A.x = 0 . In other words, if A.x = 0
, then the vector x must be zero - have 0s for all its elements.

A.x = 0

(iv) The rows of A are linearly independent.AAnd ..

(v) the columns of A are linearly independent. Where (iv) and (v) say that we
cannot get one
Aof
.x the
= 0rows or columns by making some combination of the existing
ones.
2. A matrix A which is square and symmetric
T
( A = A T , whereAA
.
x
=is0the transpose of A; i.e. a ij = a ji ) is said to be positive
T
definite if: x .A.x > 0
3. If the matrixAA=isGpositive
definite then it can be decomposed in exactly one
.G T
way into a product: A = G.G T
such that matrix G is lower triangular and has positive entries on the main diagonal.
G decomposition.
This is known as Cholesky

62

Numerical Methods in Reservoir Simulation

SOLUTIONS TO EXERCISES

EXERCISE 1.
Apply finite differences to the solution of the equation:

dy = 2. y 2 + 4
dt
where, at t = 0, y(t = 0) = 1. Take time steps of t = 0.001 (arbitrary time units) and
step the solution forward to t = 0.25. Use the notation yn+1 for the (unknown) y at
n+1 time level and yn for the (known) y value at the current, n, time level.
Plot the numerically calculated y as a function of t between t = 0 and t = 0.25 and
plot it against the analytical value (do the integral to find this).
Answer: is given below where the working is shown in spreadsheet CHAP6Ex1.xls.
This gives the finite difference formula, a spreadsheet implementing it and the analytical
solution for comparison.
SOLUTION 1.
Discretise the equation using the suggested notation as follows:

y n +1 y n
n 2

= 2.( y ) + 4
n +1t n
y y
n 2

= 2.( y ) + 4
t
which is rearranged
to the explicit formula:
y n +1 = y n + t (2.( y n )2 + 4)

ny+n1+1 yn n
nn 22
y = y + =t (22.(.(yy )) ++ 44)
n +du
t
y n +1 y n
n 2
y 1 y n = 1 tan n1 2 u

= 2.( y ) + 4
= 2.( y ) +4
u 2 + 2

t
1 easily
t up quite
which can be set
on a spreadsheet.
du
1n 2u
n +1
n =
2.( y ) + 4)
y 2 += y 2 + t(tan
might need the standard form of the following
To find the uanalytical
answer, you
1n +1
dy
n
n1 2
y n++1C= y n + t (2.( y n )2 + 4)
1 y
integral: y =2 y + t2(2=.( y ) +tan
4
dt
t
=
=
)
2
2 ydu+ ( 2 1)
2 2
1
dy= tan 11 u 1 y
2 u2y+2 + 2( 2)2 = 2 2 tan
2 = dt = t + Cdu
1
u
du
1
u

1
= tan 1
2
2

=
tan

u=2 +2 2
u +

Using this1standard
form
gives:
= 2 2 dy 2 = 1 tan 1 y = dt = t + C
2
2 1 y + (1 2 )y 2 2
1
dy
1
y
1 tandy
1+C 1 y
t
=
=
tan 1
=
2
2

=
tan
dt
t
+
C
=
=

22 2y 2 + ( 2 )22 2 2

(
)
y
+
2
2
2
2
2

2
1
1 y
2 =2 tan
2 2 = t +C
where C isy(the
constant of integration and we identify = 2 above. Therefore,
t=) = 2 2 tan 2 2 (t + C )

[
]
y2 2 (t + C )
y(1t ) =tan2 tan
[ = t +C ]
2
2 2

this becomes:

1 y1 1
C 1= tan 1tan
= t+=C0.2176049
2 2 2Heriot-Watt

Institute of Petroleum
University
2 22 1Engineering,
1 1
C
=
tan
=
0
.
2176049
y(t ) = 2 tan 2 2 (t + C )
2 2
2

1
2 2

tan 1

y
= t +C
2
63

u + = tan
du
1
u
u + = tan
du
1
u
1
1 u + dy= tan
tan
=
( 2)
21 y +dy
2 12
=
tan
( 2)
21 y +dy
2 12
2 = y 2+ ( 2 ) = 2 2 tan
2

1
1

y
= dt = t + C
y2
= dt = t + C

2
y
= dt = t + C
2

= 2
1= tan
2 1 y = t +C
y2
2 12

tan 1
= t +C

2
2 12

1 yto:
which easily rearranges
= t +C
y2(t )2=tan2 tan
2
2 2 (t + C )
y(t ) = 2 tan 2 2 (t + C )

[
]
[
]
y(t=) = 1 2 tan
tan[2 12 (t += C0).2176049
]
C
conditions since y(t = 0) = 1; it is easy to see that:
We now find C from the initial
1

tan 1
2 12

C=
tan 1
2 2

C=

2 12

12
= 0.2176049
2
1
= 0.2176049
2

which is then used in the analytic solution for y(t) above.


This is implemented in the spreadsheet CHAP6Ex1.xls. Note that at t > 0.3, both
the analytical and numerical solutions go a bit strange (very large and they start to
disagree) since the tan function has a singularity y(t) .
EXERCISE 2.
Fill in the above table using the algorithm:

Pi n +1 = Pi n +

t
( Pi n+1 + Pi n1 2 Pi n )
x 2

Hint: make up a spread sheet as above and set the first unknown block (shown grey
shaded in table above) with the above formula. Copy this and paste it into all of the
cells in the entire unknown area (surrounded by red border above).
SOLUTION 2.
If you get stuck, look at spreadsheet CHAP6Ex2.xls on the disk.
EXERCISE 3.
Experiment with the spreadsheet in CHAP6Ex2.xls to examine the effects of the
three quantities above - t, x (or NX) and the solution as t .

64

Numerical Methods in Reservoir Simulation

EXERCISE 4.
Solve the following simple example of an upper triangular matrix:

4 x1 - 14xx21 - 21xx32 + 21xx43


2 x 22xx2 + 21xx3
2

3 x3

+
+
++

11xx4 =+ 5 1x5 = 5
5
12xx4 = +122 x5 = 12
5

+ 34xx34 ++ 41xx54 =+301x5 = 30


2 x 21xx4 = 3 1x5 = 3
4

3 x5 = 153 x5 = 15

SOLUTION 4.

Answer Answer
x1 = ..........;
x2 = ..........;
x3 = ..........;
x4 = ..........;
x5 = ..........
x1 = ..........;
x2 = ..........;
x3 = ..........;
x4 = ..........;
x5 = ..........

EXERCISE 5.
fill in the table below for 10 iterations using a calculator or a spreadsheet:
POINT ITERATIVE SOLUTION OF LINEAR EQUATIONS
Iteration
counter

First guess = 0
1
2
3
4
5
6
7
8
9
10

x at iter.
x1
x2
2.0
2.0

x3
2.0

x4
2.0

SOLUTION 5.

Iteration
counter x at iter.

x1
x2
x3
x4
Answer First
- Exercise
CHAP6Ex5.xls
1
guess =5: 0see spreadsheet
2.0
2.0
2.0 - Sheet2.0
1
4.309677 1.97619
3.6925
2.784615
2
4.008926
1.411795EQUATIONS
3.498943 2.436331
POINT ITERATIVE SOLUTION
OF LINEAR
3
3.993119 1.505683 3.497206 2.503112
4
4.000937 1.500783 3.500617 2.500584
5
3.999963 1.499755 3.499965 2.499832
6
3.999986 1.500026 3.499995 2.500017
7
4.000003 1.5
3.500002 2.500001
8
4
1.499999 3.5
2.5
Converged 9
4
1.5
3.5
2.5
10
4
1.5
3.5
2.5
Note - converges fully at k = 9 iterations.

Institute of Petroleum Engineering, Heriot-Watt University

65

1
2
3
4
5
6
7
8
9
10

Iteration
counter

First guess = 0
1
2
3
4
5
6
7
8
Converged 9
10

x at iter.
x1
2.0
4.309677
4.008926
3.993119
4.000937
3.999963
3.999986
4.000003
4
4
4

x2
2.0
1.97619
1.411795
1.505683
1.500783
1.499755
1.500026
1.5
1.499999
1.5
1.5

x3
2.0
3.6925
3.498943
3.497206
3.500617
3.499965
3.499995
3.500002
3.5
3.5
3.5

x4
2.0
2.784615
2.436331
2.503112
2.500584
2.499832
2.500017
2.500001
2.5
2.5
2.5

Note - converges fully at k = 9 iterations.

EXERCISE 6.
Which is best method (i.e. that requiring the lowest amount of computational work),
BAND or LSOR, for the following problems?
(i) NX = 5,

NY = 3,

Niter = 50 (a small problem)

(ii) NX = 20, NY = 5,

Niter = 50

(iii) NX = 100, NY = 20, Niter = 70


(iv) NX = 400, NY = 100, Niter = 150
NX

NY

Niter

5
20
100
400

3
5
20
100

50
50
70
150

BAND
WB

LSOR,
WLSOR

Comment

SOLUTION 6.

66

NX

NY

Niter

50

BAND
WB
135

20

50

2500

100

20

70

8x105

400

100

150

4x108

LSOR,
WLSOR
750

Comment

For a very small problem like this, a


direct method is usually better,
although both would take very little
time even on a PC.
5000
Again a direct method is somewhat
better for a small problem.
1.4x105 As the problem grows in size,
iterative methods start to overtake
direct method.
6x106
For very large problems, iterative
methods virtually always win over
direct methods by more than an
order of magnitude.

Numerical Methods in Reservoir Simulation

EXERCISE 7.
The linear equations which arise in reservoir simulation may be solved by a direct
solution method or an iterative solution method. Fill in the table below:

Direct solution method

Iterative solution method

1.

1.

2.

2.

1.

1.

2.

2.

Give a very
brief description
of each method

Main advantages

Main
disadvantages

Institute of Petroleum Engineering, Heriot-Watt University

67

68

Numerical Methods in Reservoir Simulation

Institute of Petroleum Engineering, Heriot-Watt University

69

Permeability Upscaling

CONTENTS
1

SINGLE-PHASE FLOW
1.1
INTRODUCTION
1.2
Upscaling Porosity and Water Saturation
1.3
Averaging Permeability
1.3.1 Flow Parallel to Uniform Layers
1.3.2 Flow Across Uniform Layers
1.3.3 Flow through Correlated Random Fields
1.3.4 Additional Averaging Methods
1.3.5 Summary of Permeability Averaging
1.4
Numerical Methods
1.4.1 Recap on Flow Simulation
1.4.2 Boundary Conditions
1.5
Upscaling Errors
1.5.1 Correlated Random Fields
1.5.2 Evaluating the Accuracy of Upscaling
1.5.3 Upscaling of a Sand/Shale Model
1.6
Summary of Single-Phase Upscaling

TWO-PHASE FLOW
2.1
Introduction
2.2
Applying Single-Phase Upscaling to a
Two-Phase Problem
2.3
Improving Single-Phase Upscaling
2.3.1 Non-Uniform Upscaling
2.3.2 Well Drive Upscaling
2.4
Introduction to Two-Phase Upscaling
2.5
Steady-State Methods
2.5.1 Capillary-Equilibrium
2.6
Dynamic Methods
2.6.1 Introduction
2.6.2 The Kyte and Berry Method
2.6.3 Discussion on Numerical Dispersion
2.6.4 Disadvantages of the Kyte and Berry
Method
2.6.5 Alternative Methods
2.6.6 Example of the PVW Method
2.7
Summary of Two-Phase Flow

ADDITIONAL TOPICS
3.1
Upscaling at Wells
3.2
Permeability Tensors
3.2.1 Flow Through Tilted Layers
3.2.2 Simulation with Full Permeability
Tensors
3.3
Small-Scale Heterogeneity
3.3.1 The Geopseudo Method
3.3.2 Capillary-Dominated Flow

3.3.3 Geopseudo Example


3.3.4 When to use the Geopseudo Method
3.4 Uncertainty and Upscaling
3.5 Upscaling Summary
4

REFERENCES

Learning Objectives

After reading through this Chapter, the Student should be able to do


the following:

Appreciate why upscaling is necessary.

Know how to calculate effective permeability in simple models by


averaging.

Understand how to perform numerical upscaling of single-phase flow.

Be aware of the effects of heterogeneity on two-phase flow.

Realise the limitations of applying single-phase upscaling to a two-phase


problem.

Know how to carry out steady-state, capillary-equilibrium upscaling for twophase flow.

Become familiar with two-phase dynamic upscaling (the Kyte and Berry
Method), and understand the advantages and disadvantages of applying dynamic
upscaling.

Understand how to upscale around a well.

Appreciate that permeability is a full tensor property.

Know how to upscale from the core-scale to the scale of a geological model,
taking account of fine-scale structure and capillary effects.

Permeability Upscaling

1 SINGLE-PHASE FLOW
1.1 Introduction

Reservoir modelling often involves generating multi-million cell models, which


are too large for carrying out flow simulations using conventional techniques. The
number of cells must therefore be reduced by upscaling (Figure 1). Some quantities, such as porosity and water saturation, are easy to upscale, because they may
be averaged arithmetically. However, other quantities notably permeability are
much more difficult to upscale.
Full-field model

Geological model

Figure 1
Upscaling Example

We usually refer to the upscaled permeability as the effective permeability. The


effective permeability is defined as the permeability of a single homogeneous cell
which gives rise to the same flow as the fine-scale distribution when the same pressure
gradient is applied. We assume that Darcys Law holds at the coarse scale:

Q=

Ak eff P
x

(1)

where Q = total flow, A = area, keff = effective permeability, = viscosity, and P/x
is the pressure gradient.
True effective permeability is an intrinsic property of the model and ought to be
independent of the applied boundary conditions (Section 1.4). However, in practice,
the effective permeability often does depend on the boundary conditions, and on the
method used for calculation. Upscaling must always be carried out with care in order
to obtain sensible results.
In Figure 1, the geological model on the left is a fine-scale model with 20 million cells,
and the coarse-scale model on the right consists of about 300,000 cells. Each of the
coarse-scale cells contains an effective permeability. An example of fine-scale and
coarse-scale grids is shown in the 2D model in Figure 2. An effective permeability
is calculated for each coarse-scale cell, either by averaging the fine-grid values, or
by performing a numerical simulation.

Institute of Petroleum Engineering, Heriot-Watt University

fine grid

coarse grid
one coarse cell

Figure 2
The upscaling procedure

In this section of the upscaling course, we assume that there is only one phase present
water or oil, and that we have steady-state linear flow. We show how simple averaging may sometimes be used to estimate upscaled parameters, and then move on to
methods which involve numerical simulation. This is followed by a set of examples
which demonstrate how errors may arise, and how to avoid them.
1.2 Upscaling Porosity and Water Saturation

We start by averaging porosity and water saturation, using a simple model (Figure 3).
(Note that the water saturation is not required for a single-phase problem. However,
we include it here because it is simple to upscale.) There are 10 grid blocks of size
1 m3, 4 of which have a porosity of 0.15, and 6 of which have a porosity of 0.20.
= 0.15

= 0.20

Sw = 0.50

Sw = 0.40

Figure 3
Example for averaging
porosity and water
saturation

The average porosity, , is given by:

total pore volume


total volume

(2)

Therefore, in this case:

4 0.15 + 6 0.20
= 0.18.
10

When averaging the water saturation, we need to take the porosity into account. In the
previous example, suppose the water saturation was 0.5 in the blocks with porosity of
0.15, and 0.4 in the blocks with porosity 0.2, then the average water saturation is:

Sw =

total amount of water


total pore volume

Here, the average water saturation is:

(3)

Permeability Upscaling

4 0.15 0.5 + 6 0.20 0.4


4 0.15 + 6 0.20
0.3 + 0.48 0.78
=
= 0.433.
=
1.8
1.8

Sw =

1.3 Averaging Permeability

In some simple models, such as parallel layers or a random distribution, the effective
permeability may be calculated by averaging.
1.3.1 Flow Parallel to Uniform Layers

P1

P2

Qi

ki, ti

Figure 4
Along-layer flow

x
Consider a set of (infinite) parallel layers of thickness, ti and permeability ki, where i
= 1, 2, .. n (the number of layers). The effective permeability of these layers is given
by the arithmetic average, ka.
n

k eff = k a =

t k
i =1
n

t
i =1

(4)

(Equation (4) may be proved by applying a fixed pressure gradient along the layers.)
Example 1
x
Figure 5
A simple, two-layer model

t1 = 3 mm, k1 = 10 mD

t2 = 5 m
mm, k2 = 100 mD

Suppose we have two layers as shown in Figure 5. The effective permeability for
flow in the x-direction is given by Equation (4), and is:

ka =

3 10 + 5 100 530
=
= 66.25 m
mD
3+5
8

Institute of Petroleum Engineering, Heriot-Watt University

1.3.2 Flow Across Uniform Layers

ki, ti

Pi

Figure 6
Across-layer flow

x
For flow perpendicular to the layers, the effective permeability is given by the harmonic average, kh:
n

k eff = k h =

t
i =1
n

ti

i =1 k i

.
(5)

(Equation (5) may be proved by assuming a constant flow rate through each layer.)
Example 2
Equation (5) may be used to calculate the effective permeability for flow across the
two layers in the model shown in Figure 5, i.e. flow in the z-direction.

kh =

8
3+5
=
= 22.86 mD
3 10 + 5 100 0.35

From Examples 1 and 2, we see that the permeability is different in different directions. In reservoirs with approximately horizontal layers, the arithmetic average
may be used for calculating the effective permeability in the horizontal direction,
and the harmonic average may be used for calculating the effective permeability in
the vertical direction.
1.3.3 Flow through Correlated Random Fields

Figure 7 shows an example of a correlated random permeability distribution. Correlated random fields are described in Section 1.5.1. Basically, correlated means that
areas of high or low permeability tend to be clustered, so that the spatial distribution
is smoother than a totally random one. The correlation length is approximately the
size of patches of high or low permeability. The longer the correlation length, the
longer will be the range of the semi-variogram for the permeability distribution.
Assuming that we are averaging over many correlation lengths, permeability should
be isotropic (same in the x-, y- and z-directions). The effective permeability for a
random permeability distribution is proportional to the geometric average, which is
given by:

Permeability Upscaling

ln( k i )

k g = exp i =1

(6)

where i = 1, 2, .. n is the number of cells in the distribution.

Correlation Length

Figure 7
A correlated, random
permeability distribution
(white = high permeability,
dark = low permeability)

The results given below have been derived theoretically for log-normal distributions,
with a standard deviation of Y, where Y = ln(k). The results depend on the number
of dimensions:

k eff = k g (1 2Y 2)

in 1D

k eff = k g

in 2 D

k eff = k g (1 + 2Y 6)

in 3D

(7)

These formulae are approximate, and assume Y is small (< 0.5). (You are not required
to know the proof.) The 1D result is an approximation of the harmonic average.
Note that the results do not depend on the correlation length of the field, provided it
is much smaller than the system size.
Also note that ka > kg > kh, and the effective permeability always lies between the
two extremes: ka and kh.
Example 3

Figure 8
A random arrangement
of the permeabilties in the
simple example

75 cells of 10 mD 125 cells of 100mD

Institute of Petroleum Engineering, Heriot-Watt University

Suppose that the permeability values in the simple example are jumbled up, so that
there are 75 small cells of 10 mD, and 125 small cells of 100 mD. See Figure 8. The
effective permeability of this model is:

k eff = k g = 10
= 10

75 log(10 ) +125 log(100 )

200

75 + 250
200

= 101.625 = 42.17 mD.


1.3.4 Additional Averaging Methods

Since averaging is very quick (compared with numerical simulation), many engineers use this technique in more complex models. Sometimes engineers increase the
accuracy by using power averaging. The power average is defined as:
1/

n
ki
k p = i =1
n

,
(8)

where is the power. The value of the power depends on the type of model, and
must be calibrated against numerical simulation (Section 1.4).
Also, sometimes, engineers use a combination of the arithmetic and harmonic averages, e.g. they take the arithmetic average of the permeabilities in each column and
then calculate the harmonic average of the columns.
1.3.5 Summary of Permeability Averaging

To summarise, there are two types of simple model in which we can calculate the
effective permeability by averaging:
Parallel layers
Correlated random fields
Since averaging is very quick, it is frequently used as an approximation for the effective permeability in more complex models.
1.4 Numerical Methods

In general, the permeability distribution will not be simple enough for us to be able
to calculate the effective permeability analytically (i.e. by averaging), and we will
have to perform a numerical simulation. We can use a finite difference method to
calculate the pressures.
1.4.1 Recap on Flow Simulation

The continuity equation tells us that there is no net accumulation or loss of fluid
within a grid block:

Permeability Upscaling

q xin + q zin = q xout + q zout .

(9)

(We are assuming incompressible rock and fluids, here.)


qzin

i,j-1

qxin

i-1,j

qxout

Figure 9
Recap on numerical flow
simulation

i,j

i+1,j

i,j+1
qzout

Darcys law is used to express the flows in terms of the pressures and permeabilities.
For example, if the grid blocks in Figure 9 are of length x and height z (and unit
width in the y-direction), then:

q xin =

k x, i 1/ 2, j z Pi , j Pi 1, j

)
(10)

where kx,i-1/2,j is the harmonic average of the permeabilities in the x-direction in blocks
(i-1,j) and (i,j). (You now should know now why the harmonic average is used here.)
The other flows are calculated in a similar manner.
It is useful to use the transmissibilities, Tx = kxz/x and Tz = kzx/z. (Assume
the width, y = 1.) We can therefore derive the pressure equation:

(T

x , i 1 / 2 , j

+ Tx, i +1/ 2, j + Tz, i , j 1/ 2 + Tz, i , j +1/ 2 Pi , j

Tx, i 1/ 2, j Pi 1, j Tx, i +1/ 2, j Pi +1, j


Tz, i , j 1/ 2 Pi , j 1 Tz, i , j +1/ 2 Pi , j +1
=0

(11)

An equation is set up for each Pij, i = 1, 2, .. nx and j = 1, 2, .. nz. The transmissibilities


are known, and using the appropriate boundary conditions, we can solve this set of
linear equations to obtain the pressure in each grid block. The effective permeability is
than calculated from the total flow and the total pressure drop, as described below.
Note that the boundary conditions are applied to each coarse grid cell in turn, and
they may not be a good approximation to the pressures which would arise in a finegrid simluation. This leads to errors in the results. Upscaling errors are discussed
in Section 1.5.
1.4.2 Boundary Conditions

Boundary conditions are required to specify what happens at the edges of the
model.

Institute of Petroleum Engineering, Heriot-Watt University

a) No-Flow Boundaries
no flow through the sides

P1

P2
Figure 10
Fixed pressure, or no-flow,
boundary conditions

no flow through the sides

The pressure is fixed on two sides of the model, and no flow is allowed through the
others sides of the model. This type of boundary condition is suitable for models
where there is little cross-flow: for example, models with approximately horizontal
layers, or a random distribution. These are the most commonly applied boundary
conditions. Figure 11 illustrates how an effective permeability may be calculated
in the x-direction.
Pressure= P1
on left face

Pressure= P2
on right face

y
x

Area, A

Flow Rate, Q

Figure 11
The calculation of effective
permeability using no-flow
boundary conditions

1. Solve the steady-state equation to give the pressures, Pij, in each grid block.
2. Calculate the inter-block flows in the x-direction using Darcys Law. (See
Equation 10.)
3. Calculate the total flow, Q, by adding the individual flows between two y-z
planes. (Any two planes will do, because the total flow is constant.)
4. Calculate the effective permeability for flow in the x-direction, using the
equation:

Q=

k eff , x A( P1 P 2)
L

(12)

Repeat the calculation for flow in the y- and z-directions, to obtain keff,y and keff,z.
(b) Periodic Boundary Conditions
Periodic boundary conditions are useful for calculating the effective permeabilities
in models where there are infinitely repeated geological structures in each direction.
(See Section 3.2.) The use of periodic boundary conditions ensures that we also

10

Permeability Upscaling

have an infinitely repeated pattern of flows and pressure gradients. In the example
shown in Figure 12, there is a net pressure gradient in the x-direction. The blocks
are numbered i = 1, 2, ..nx in the x-direction, and j = 1, 2, .. nz in the z-direction.
x

P(i,0) = P(i,nz)

Figure 12
Periodic boundary
conditions

P(nx+1,j) = P(1,j)-P

P(0,j) = P(nx,j)+P

P(i,nz+1) = P(i,1)

One advantage of using periodic boundary conditions, is that fluid can flow through
the sides of the model. This method can be used to calculate a full tensor effective
permeability (Section 3.2).
(c) Linear Pressure Boundary Conditions
In linear pressure boundary conditions (Figure 13), the pressure is fixed at each end,
as in the fixed-pressure boundary conditions. Then, the pressure at the edges of the
model is interpolated linearly from one side to the other. Like the periodic boundary
conditions, the linear pressure boundary conditions allow flow through the edges.
P1

P2

P1

Figure 13
Linear pressure boundary
conditions

P2

P1

P2

(d) Flow Jacket, or Skin


To reduce the effect of boundary conditions when calculating the effective permeability,
some engineers perform the simulations on a larger grid than necessary. The extra
grid blocks round the edges are referred to as a jacket or skin. See Figure 14.

Institute of Petroleum Engineering, Heriot-Watt University

11

boundary conditions applied


to outer edges of model

keff calculated for


this block

Figure 14
Example of a flow jacket
round a model. In this case
the jacket is 4 cells thick

1.5 Upscaling Errors

In the process of upscaling, information about the fine-scale structure is lost, and
upscaling usually gives rise to errors. However, in some cases, the errors will be
larger than in others. In this section, we examine a series of models which show
examples of where upscaling is successful and where it is not.
1.5.1 Correlated Random Fields

We introduce the concept of a correlated random field. Although the permeability


distribution in real rocks may not follow this type of model, it is a useful way to
parameterise heterogeneity. We assume that the probability density function (pdf)
of the model is normal or log normal, as shown in Figure 15. In an isotropic model
(i.e. same in all directions), the field is then characterised by three parameters: the
mean, , the standard deviation, , and correlation length, . The standard deviation
determines the width of the pdf (i.e. the permeability contrast), and the correlation
length determines approximately the distance over which the permeability values are
similar. Figure 16 shows examples of the 4 models with varying and .
0.06

a)

0.06
0.05
Frequency

Frequency

0.05
0.04
0.03
0.02

0.03
0.02
0.01

0.00
40

0.00
0.5

60

80
100 120 140
Perm eabil ity (m D)

160

c)

0.05
Frequency

0.04

0.01

0.06
0.04
0.03
0.02
0.01
0.00

12

b)

500
1000
1500
Per m eab ility (m D)

2000

1.0

1.5
2.0
2.5
3.0
log (p erm eabil ity)

3.5

Figure 15
Normal and log-normal
permeability distributions.
a) Normal distribution
with mean = 100 mD, and
standard deviation = 20
mD.
b) Log-normal distribution
with mean = 2.0, and
standard deviation = 0.5.
c) Log-normal distribution
as above, but with
permeability plotted on
the x-axis, rather than
log(permeability)

Permeability Upscaling

Figure 16
Models with different
standard deviations and
correlations lengths

a) small , small

b) large , small

a) small , large

b) large , large

Permeability (mD)
0

50

100

150

200

1.5.2 Evaluating the Accuracy of Upscaling

One way to evaluate the accuracy of upscaling is to compare upscaling in two stages,

k 2eff , with upscaling in a single stage, k1eff , as shown in Figure 17. If upscaling is
2
1
accurate, then k eff = k eff . This is the case for upscaling along and across parallel
layers in the idealised models of Sections 1.3.1 and 1.3.2.
fine-scale model

k1eff

single cell

k2eff

Figure 17
Comparison of one-stage
and two-stage upscaling

The accuracy of scale-up is affected by the correlation length and standard deviation
of the distribution, and we use the method shown in Figures 10 and 11 to demonstrate
this effect. Instead of generating many fine-scale models with different correlation
lengths, we create 1 fine-scale model, but upscale by different factors so that the

Institute of Petroleum Engineering, Heriot-Watt University

13

coarse block size varies relative to the correlation length. Figure 18a shows a finescale model with a correlated random permeability distribution. The model has 400
x 400 grid cells, each of size 1 m3. The permeability distribution is ln-normal, with
a mean of 4.6 (corresponding to 100 mD), and a correlation length of 10 m. Three
different versions of the model were created with different standard deviations: 0.5,
0.75 and 1.0. The following scale-up factors were tested:

4 4, 5 5, 8 8, 10 10, 16 16, 20 20, 40 40, 80 80


In terms of the correlation length, this gives coarse-scale cells of size:
0.4, 0.5, 0.8, 1.0, 1.6, 2.0, 4.0, 8.0.
Figure 18b and c show examples of coarse-scale models with scale-up factors of 5
and 50. In each case the ratio of two-stage upscaling to single-stage upscaling was
calculated. The results are plotted in Figure 19. The results are least accurate when
the scale-up factor is 10 50, i.e. when the coarse block size is 1 5 times the correlation length. Also, the error increases with the standard deviation of the model,
as one might expect.
a)

b)

c)

Permeability (mD)
0.1

10

100

1000

10000

Figure 18
a) Fine-scale model with
400 x 400 cells; b) coarsescale model with 80 x 80
cells; c) coarse-scale model
with 8 x 8 cells

keff2/keff1

1.03

1.02

sigma = 1.00
sigma = 0.75
sigma = 0.50

1.01
1.
01

1.00
0

20

40

60

80

Scale-up Factor

The conclusions from these examples are:

14

Upscaling will be least accurate when the coarse cell size is comparable to, or
slightly larger than the correlation length.
Upscaling errors increase as the standard deviation of the model increases.

Figure 19
2
1
The ratio of k eff k eff for
different scale-up factors

Permeability Upscaling

1.5.3 Upscaling of a Sand/Shale Model

Upscaling errors are largest in models where there are high permeability contrasts.
Unfortunately, high permeability contrasts frequently occur in reservoir rocks. For
example, we often require to model the following:
Low permeability shales in a high permeability sandstone
Low permeability faults in a high permeability sandstone
High permeability channels in a low net/gross reservoir
High permeability fractures in a low permeability reservoir
All these cases are difficult to model. As an example, we consider a sand/shale model,
where the shale has zero permeability. Figure 20 shows the fine-scale model, which
has to be upscaled to 3 coarse blocks, as shown. Since there is a shale lying across
each coarse block, each coarse block will have zero permeability in the z-direction
(vertical). However, fluid can flow through the model vertically, as shown. This
error arises because the coarse block size is similar to the characteristic length of the
shales. Upscaling would be more accurate, if the coarse block size was much larger,
or much smaller than the shales. Alternatively, using a skin or flow jacket will
increase the accuracy of upscaling (Section 1.4.2, Figure 14).

Figure 20
Sand/shale model
1.6 Summary of Single-Phase Upscaling

The main points to remember from this section are:

Fine-scale geological models usually require upscaling for full-field


simulation.
Upscaled permeability is generally referred to as effective permeability.
Some quantities, such are porosity and water saturation are easy to upscale,
because they may be averaged arithmetically.
In some simple models, permeability may also be upscaled using averaging, as
follows:
the arithmetic average for along-layer flow;
the harmonic average for across-layer flow;
the geometric average for a random distribution.
In more complex models, the effective permeability is calculated using a
numerical simulation.
Different boundary conditions may be used when calculating the effective
permeability numerically: constant pressure (or no-flow), periodic and linear.
Upscaling is least accurate when the coarse cell size is comparable to, or slightly
larger that the correlation of the permeability distribution.
Upscaling errors increase as the standard deviation of the model increases.

Institute of Petroleum Engineering, Heriot-Watt University

15

2 TWO-PHASE FLOW
2.1 Introduction

Often we need to simulate two-phase systems, e.g. a water flood or a gas flood of an
oil reservoir, or an oil reservoir with a gas cap or an aquifer. The aim of upscaling
in this case is to calculate a coarse-scale model which can reproduce the flow rates
of the different fluids. The coarse model should also provide a good approximation
to the saturation distribution in the reservoir with time.
The paths which the injected fluid takes through the reservoir depends on the forces
present:
Viscous due to injection of a fluid
Capillary
Gravity
Therefore, the balance of forces should be taken into account during upscaling.
Before learning how to upscale two-phase flow, we show the effects which geological heterogeneity may have on hydrocarbon recovery.
Consider the following simple model (Figure 21), with alternating horizontal layers of
100 mD and 10 mD (referred to as facies 1 and facies 2). We assume that the model
is filled with oil and connate water initially, and simulate a water flood, by injecting
at uniform rate at the left side, and producing from the right side (at constant bottomhole pressure). The density of the two fluids is the same for this example, so that
there are no gravity effects. Figure 22 shows the relative permeabilities and capillary
pressures, and Table 1 lists the properties of the 3 cases simulated with this model.
100 mD
10 mD

12

0.9

krw 1
kro 1
krw 2
kro 2

8
6

Pc 1
Pc 2

0.8
0.7
0.6

Rel Perm

Cap Pressure

10

0.5
0.4
0.3

0.2

0.1
0

0
0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.2

0.3

Case
1
2
3

Porosity
facies 1
facies 2
0.2
0.2
0.2
0.05
0.2
0.1

0.4

0.5

0.6

Water Saturation

Water Saturation

16

Figure 21
Simple layered model for
demonstrating viscous and
capillary effects

Rel Perm/Pc Curve No.


facies 1 facies 2
1
1
1
1
1
2

0.7

0.8

Figure 22
The relative permeability
and capillary pressure
curves

Flow Regime
viscous
viscous
visc+capillary

Table 1
Properties of the first set of
examples

Permeability Upscaling

Figure 23 shows the distribution of oil saturation after the injection of 0.2 PV, for
Cases 1 and 2. Both cases use only a single relative permeability curve and the flow
regime is viscous-dominated. Water flows faster along the high permeability layers,
as one would expect. However, notice that this effect is reduced when the porosity
of facies 2 is reduced.

Figure 23
The oil saturation for Cases
1 and 2, after the injection
of 0.2 PV

Oil Saturation

0.3

0.4

0.5

0.6

0.7

In Case 3, both relative permeability and capillary pressure tables are used. These
curves are typical of a water-wet rock: the capillary pressure is much higher in the
low permeability facies, and the connate water saturation is higher. In this case, water
is imbibed along the low permeability layers, and also there is cross-flow from the
high permeability layers to the low permeability ones (Figure 24). Due to the effects
of capillary pressure, the front is nearly level in the two facies.

Figure 24
The oil saturation for Case
3, after the injection of 0.2
PV

Oil Saturation

0.3

0.4

0.5

0.6

0.7

Figure 25 shows the recovery as a function of pore-volumes injected, and demonstrates that models may have the same effective absolute permeability, but different
recoveries.

Institute of Petroleum Engineering, Heriot-Watt University

17

0.5
0.4
0.3

Case 1
Case 2
Case 3

0.2
0.1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Pore Volumes Injected

Figure 25
Cumulative recovery and
watercut for Cases 1 - 3

In the next example, graded models are used, as shown in Figure 26. There are two
versions: Case 4 with permeability increasing upwards (referred to as coarsening-up)
and Case 5 with permeability decreasing upwards (referred to as fining-up). The
permeabilities range from 200 mD to 1000 mD, the porosity was kept constant at
0.2, and the first relative permeability curve was used. In this model, however, the
densities of the fluids were different: the density of water was set to 1000 kg/m3 and
that of oil was set to 200 kg/m3.
a) Coarsening-up

b) Fining-up

1000 mD

200 mD

200 mD

Figure 26
The graded layer models for
Cases 4 and 5

1000 mD

Again a waterflood was performed, and the results are shown in Figure 27. Since
water is more dense than oil, water has a tendency to slump down. In Case 4 (coarsening-up), this tendency is reduced by the fact that the viscous forces tend to move
the fluid faster in the upper layers. However, in Case 5 (fining-up), the slumping
effect is reinforced by the viscous force moving fluid faster along the lower layers.
This means that the breakthrough time is earlier in Case 5 than in Case 4, as shown
in Figure 27. The effective absolute permeability in these two models is the same,
but the two-phase flow effects are different, due to the effect of gravity. (If the density of water equalled the density of oil, the recovery and watercut curves would be
identical.)
a) Coarsening-up

a) Fining-up

Oil Saturation

0.3

18

0.4

0.5

0.6

0.7

Figure 27
The oil saturation after the
injection of 0.2 PV in the
graded layer models

Figure 28
Cumulative recovery and
watercut for the graded
layer models

0.5

1.0

0.4

0.8

Water Cut

Fractional Recovery

Permeability Upscaling

0.3
0.2
0.1
0.0
0.0

0.6

Case 4
Case 5

0.4
0.2

0.1

0.2

0.3

0.4

0.5

0.0

0.6

0.0

0.1

Pore Volumes Injected

0.2

0.3

0.4

0.5

0.6

Pore Volumes Injected

In summary, in a viscous-dominated flood, permeability heterogeneity disperses the


flood front (Cases 1 and 2), so that breakthrough occurs earlier, and the water cut
curve rises less steeply. However, the effect of heterogeneity also depends on the
balance of fluid forces. In a water-wet system, capillary pressure can help the front
to advance more evenly along the layers (Case 3). (More information on capillary
effects is given in Section 3.3.) Gravity effects may increase or reduce the viscous
effects, depending on permeabilities in the model (Cases 4 and 5).
2.2 Applying Single-Phase Upscaling to a Two-Phase Problem

Most engineers only perform single-phase upscaling although, as shown above, heterogeneities give rise to a variety of effects in two-phase flow. The reason for this
is that two-phase upscaling is time-consuming and the results are not always robust
(i.e. they may contain large errors). We deal with two-phase upscaling in Sections
2.4 2.6.
Figure 29 (left side) shows a very heterogeneous 2D model, with correlated random
permeabilities. The permeability distribution was ln-normal (i.e. natural logs), with
a standard deviation of 2.0 The model is assumed to be in the horizontal plain. The
details of the model are given in Table 2. The model was upscaled using the pressure solution method with no-flow boundaries. Three different scale-up factors were
used, and the coarse-scale models 1 and 3 are also shown in Figure 29 (middle and
right side).

Figure 29
Fine- and coarsescale models used for
demonstrating the effects
of applying single-phase
upscaling to a two-phase
problem

Table 2

log10(k)

-2

Model
Fine
Coarse 1
Coarse 2
Coarse 3

-1

Number of Cells
105 x 105
21 x 21
15 x 15
7x7

Cell dimension (m)


5
25
35
75

Institute of Petroleum Engineering, Heriot-Watt University

Scale-up Factor
5x5
7x7
15 x 15

19

A waterflood with a quarter 5-spot well pattern was performed (i.e. 2 wells in
diagonally opposite corners). The same relative permeability curve was used for
the whole model (Figure 30), and the capillary pressure was set to zero. The flood
was therefore viscous-dominated. The viscosity of water was 0.3 and that of oil was
3.0. The resulting recovery curves are shown in Figure 31. As one might expect, the
error increases with the scale-up factor.

Relative Permeability

1
0.8
0.6

krw
kro

0.4
0.2
0
0

0.2

0.4

0.6

0.8

Water Saturation

Figure 30
The relative permeability
curve used for the random
model. (Capillary pressure
was set to zero)

0.6

Fractional Recovery

0.5
0.4

fine
ups 5x5
ups 7x7
ups 15x15

0.3
0.2
0.1

Figure 31
Comparison of recovery for
different scale-up factors

0
0

0.2

0.4

0.6

0.8

Pore Volumes Injected

The model was modified by reducing the standard deviation to 0.2 (lowering the
permeability contrast), and the simulations were repeated with the low heterogeneity
model. The recovery is shown in Figure 32 for the scale-up factor of 15 x 15. As
expected, the errors are smaller for the low heterogeneity model.

Fractional Recovery

0.6
0.5
0.4

lo het fine
lo het coarse
hi het fine
hi het coarse

0.3
0.2
0.1
0

0.2

0.4

0.6

0.8

Pore Volumes Injected

In addition, the errors caused by using only single-phase upscaling, are larger when
the coarse block size similar to the correlation length. Also, the upscaling error tends
20

Figure 32
Comparison of recovery for
models with different levels
of heterogeneity

Permeability Upscaling

to be larger in unstable floods (injected fluid is of lower viscosity than the in situ
fluid) than in stable floods.
In summary, single-phase upscaling may be adequate for upscaling two-phase systems, provided that:

The scale-up factor is small


The permeability contrasts are small
The correlation length is very large, or very small compared to the coarse cell
size
The flood is stable (favourable mobility ratio)
The flood is not capillary-dominated (See Section 3.3)
The flood is not gravity-dominated

2.3 Improving Single-Phase Upscaling

There are two approaches which may make single-phase more accurate when applying
it to two-phase problems. The first is to use non-uniform upscaling, and the second
is to perform a global single-phase simulation (i.e. over the whole fine-scale model)
using the correct boundary conditions, including wells. We refer to this second
method as Well Drive Upscaling (WDU).
2.3.1 Non-Uniform Upscaling

Consider a model with horizontal layers, as shown in Figure 33. There is a high
permeability streak running across the model. The model details are given in Table 3.
In both coarse-scale models there are 3 coarse cells in the vertical direction. In model
Coarse 1, the cells are each 5 m thick. However, in model Coarse 2, the thicknesses
are: 7 m, 1 m, 7 m, so that the high permeability streak is still resolved.
b) C
Coa
oars
oa
rse
rs
e1

c) Coar
c)
Coar
oarse
se 2

Figure 33
Model with a high
permeability streak

Table 3

Model
Fine
Coarse 1
Coarse 2

No. of cells
100 x 15
20 x 3
20 x 3

Cell size (m)


1x1
5x5
variable

Figure 34 shows the recovery for these models. It can be seen that the model with
uniform coarse cells (Coarse 1) gives very inaccurate results, and the model which
maintains the high permeability streak (Coarse 2) is much more accurate.

Institute of Petroleum Engineering, Heriot-Watt University

21

Fractional Recovery

0.25
0.20
0.15

fine
coarse 1
coarse 2

0.10
0.05
0.00

0.0

0.1

0.2

0.3

Figure 34
Recovery and watercut
for fine- and coarse-scale
models

0.4

Pore Volumes Injected

In a practical upscaling application, much attention is paid to upgridding the model


i.e. deciding how to amalgamate the layers, so that upscaling is as accurate as possible. The coarse grid may also be non-uniform in the x- and y-directions.
Several methods for performing non-uniform upscaling have been developed. For
example, Durlofsky et al. (1996, 1997) first carry out a single-phase simulation. Then
they use the inter-block flows to determine the coarse block boundaries. Smaller
coarse blocks are assigned to regions where there are high flow rates.
2.3.2 Well Drive Upscaling

When upscaling using numerical simulation, boundary conditions are applied to


each coarse-scale cell in turn (Figure 2.) These are called local boundary conditions. However, the boundary conditions described in Section 1.4.2 may be quite
different from the pressures which actually occur in a fine-scale simulation, leading to inaccuracies in the upscaled model. To overcome this problem, we may use
global boundary conditions. Figure 35 demonstrates the difference between local
and global boundary conditions.
Boundary conditions
applied to coarse cell

P1

Injector

P2

Producer

Local

Global

In the Well-Drive Upscaling method (WDU), a single-phase simulation is performed


on the whole fine grid (Zhang et al, 2005). (It is feasible to perform a single pressure
solve on grids with several million grid cells.) Then the effective transmissibility
between coarse-scale cells is calculated, rather than the effective permeability. The
upscaled transmissibility is give by:

22

Figure 35
Local and global boundary
conditions

Permeability Upscaling

T=

P P

(13)

Where q denotes the fine-scale flows (Figure 36) and PI and PII are the (pore-volume
weighted) average pressures in coarse cells I and II.

Figure 36
Upscaling transmissibility

Upscaling is also performed at the wells, using the method described in Section
3.1. This method produces very accurate single-phase upscaling, which leads to an
increase in the accuracy of two-phase flow at the coarse scale.
Many tests on upscaling methods have been carried out using the model generated
for the 10th SPE comparative solution project, which was on upscaling (Christie and
Blunt, 2001). This model is referred to as the SPE 10 model (Figure 37a). We use
layer 59 (Figure 37b) as an example of well drive upscaling, because this layer is
particularly heterogeneous. There is an injection well at the centre and production
wells in each corner.
a)

Figure 37
The SPE 10 model, and
layer 59

b)

P1

P2

P4

P3
log10(k)

-3

-2

-1

Layer 59 was upscaled using the WDU method and also the conventional method
with local boundary conditions. Figure 38 shows the oil saturation distribution. It
can clearly be seen that the WDU method reproduces the results of fine-scale model
much better than the conventional approach. This is because appropriate boundary
conditions have been applied to the model.

Institute of Petroleum Engineering, Heriot-Watt University

23

fine

local

WDU

Figure 38
Comparison of oil
saturation distribution
in fine- and coarse-scale
simulations of Layer 59 of
the SPE 10 model

Soil

0.20

0.35

0.50

0.65

0.80

2.4 Introduction to Two-Phase Upscaling

So far, when performing upscaling, we have assumed that there is only one phase
present, and that the flow is in a steady state. We only need to upscale the absolute
permeability. However, when there are two phases flowing, such as water displacing oil, the system is not, in general, in a steady state. We need to simulate finescale floods in order to upscale relative permeability and capillary pressure. This
is referred to as dynamic upscaling, and the upscaled relative permeabilities are
known as pseudo relative permeabilities, or pseudos. Pseudos can be calculated to
take account of physical dispersion, and also to compensate for numerical dispersion
(Section 2.6.3).
When upscaling, we should use the phase permeabilities:

k f = k abs k rf

(14)

Where f stands for fluid oil, gas or water. Generally, we assume that both the
absolute and the relative permeabilities are homogeneous and isotropic at the smallest
scale ( k x = k z ). As we upscale, the absolute and relative permeabilities may become
anisotropic ( k rx k rz ). To obtain effective (or pseudo) relative permeabilities, the
absolute permeability must be scaled-up separately. Then the pseudo relative permeability is calculated as follows:

k rf , x = k f , x k abs, x

(15)

Similar equations are used for flow in the y- and z- directions.


2.5 Steady-State Methods

If fluids are in a steady state, the saturation does not change with time and the fractional
flow (flow of water/total flow) is constant. Although floods are dynamic processes,
sometimes a flood may approach a steady state. For example, over small scales (20
cm, or less), oil and water may come into capillary equilibrium.

24

Permeability Upscaling

In a steady-state upscaling method, we assume that within a short interval of time the
zone of interest is in a steady-state, but we allow the fluid saturation to change gradually, so that a full range of saturation is obtained. At steady-state, the water saturation
does not change with time, i.e. Sw/t = 0, so the continuity equation becomes:

u f = 0,

(16)

where u is Darcy velocity, and f is fluid. From Darcys law:

( k f Pf ) = 0.

(17)

There are several steady-state methods, depending on the balance of forces:


Capillary equilibrium,
Vertical equilibrium (gravity-dominated flood)
Viscous-dominated steady-state
We concentrate here on the capillary equilibrium method.
The advantage of steady-state methods is that they turn two-phase upscaling into a
series of single-phase upscaling calculations. This means that steady-state methods
are feasible for models with large numbers of grid cells. (See, for example, Pickup
and Stephen, 2000; and Pickup et al, 2000.)
2.5.1 Capillary-Equilibrium

Assume that the injection rate is very low, gravity forces are negligible, and that the
fluids have come into capillary equilibrium with a coarse-scale cell. This means that
the saturation distribution is determined by the capillary pressure curves.
The method is as follows:
1. Choose a Pc level.
2. Determine the water saturations, and then the relative permeabilities.
3. Calculate the pore volume-weighted average water saturation.
4. Calculate the phase permeabilities: ko = kabskro, kw = kabskrw.
5. Calculate the effective water phase permeability, kw
6. Calculate the effective oil phase permeability, ko
7. Calculate the relative permeabilities, krw = kw/kabs, etc.
8. Repeat the process with another value of Pc.
Steps 5 and 6 may be carried out analytically or numerically, depending on the
distribution.
Institute of Petroleum Engineering, Heriot-Watt University

25

Example 4
Consider a model with two layers of equal thickness, as shown in Figure 39. The
absolute permeabilities are 100 mD and 20 mD. Assume that the porosity in each
layer is equal to 0.2. The relative permeability and Pc curves for each layer are
shown in Figure 40.

kabs (mD)
100

Figure 39
Model with horizontal
layers

20

0
8
6
4

0.8

lo

Rel Perm

Cap Pressure

12

hi

2
0
0.2

0.3

0.4

0.5

0.6

Water Saturation

0.7

0.8

0.6

hi

lo

0.4

hi

0.2
0

0.2

0.3

0.4

0.5

0.6

Water Saturation

lo
0.7

0.8

Figure 40
Relative permeability and
capillary pressure curves

Using the arithmetic and harmonic averages (Section 1.3), the effective permeability
is:

k x = 60.00

k z = 33.33

Suppose we choose a capillary pressure of Pc = 0.45.


In the high perm layer: Sw = 0.34,
krw = 0.0013, kw = 0.13, kro = 0.5, ko = 50.
In the low perm layer: Sw = 0.44,
krw = 0.0016, kw = 0.032,

kro = 0.48, ko = 9.6.

Figure 41 shows the phase permeabilities.

kw (mD)

ko (mD)

0.13

50.0

0.032

9.6

Since the layers are of equal width, the average saturation is Sw = 0.39. The effective
phase permeabilities are then calculated using the arithmetic and harmonic averages.
Then the relative permeabilities are calculated using Equation 15.

26

Figure 41
Phase permeabilities

Permeability Upscaling

k wx = 0.081

k rwx = 0.081 / 60.00 = 0.00135

k wz = 0.051

k rwz = 0.051 / 33.33 = 0.00153

k ox = 29.8

k rox = 29.8 / 60.00 = 0.50

k oz = 16.1

k roz = 16.1 / 33.33 = 0.48

Note that the kv/kh ratio ( = k z k x ) is different for oil and water:

k w, z k w, x = 0.63

k o, z k o, x = 0.54

Effective relative permeability curves may be derived by repeating this calculation for
a range of capillary pressure values (Figure 42). The capillary-equilibrium method
is useful as a quick method for upscaling small-scale models (Section 3.3). However,
it is only valid in cases where the flow rate is very low.
0.9
0.8

Rel Perm

0.7

krox

0.6
0.5
0.4

kroz

0.3

krwz

0.2

Figure 42
Effective relative
permeability curves

0.1
0
0.2

0.3

0.4

0.5

0.6

Water Saturation

krwx
0.7

0.8

2.6 DYNAMIC METHODS


2.6.1 Introduction

For dynamic (or non steady-state methods), we need to perform a two-phase flow
simulation on a fine grid. There are basically two types of dynamic method:
a) Weighted Pressure Methods
As in single-phase numerical upscaling, a common approach in two-phase upscaling
is to sum the flow, average the pressure gradient and use Darcys Law to obtain the
pseudo phase permeability. However the pressure may be averaged in different ways.
Here, we shall concentrate on the Kyte and Berry (1975) method.
b) Total Mobility Methods
In total mobility methods, we avoid averaging the pressure, and scale-up the total
mobility. Then the average fractional flow is used to calculate the pseudo relative
permeabilities.
The total mobility is:

t = o + w =

k ro k rw
+
.
o w

Institute of Petroleum Engineering, Heriot-Watt University

(18)
27

The fractional flow is the flow of water divided by the total flow:

fw =

qw
q
= w
qo + qw qt

(19)

Again there are a number of variations of this method, the most commonly used
being that of Stone (1991).
2.6.2 The Kyte and Berry Method

A simple version of the Kyte and Berry (1975) method is presented here, using the
grid shown in Figure 43.
i=1

9 10

j=1
2
3
4
5

z
x

Pseudo calculated
for this coarse block

DZ
Z

DX
X

Figure 43
Model used for describing
the Kyte and Berry Method.
The thickness of the model
is y

The diagram shows two coarse grid blocks, each of which is made up of 5 x 5 fine
blocks. The equations below show how to calculate the pseudo relative permeabilities
and capillary pressure for the left coarse block.
The first step is to perform a fine-scale, two-phase simulation (e.g. in ECLIPSE),
saving the pressures and inter-block flows at specified intervals of time (in the re-start
files). The method proceeds as follows:
1. Calculate the effective absolute permeability in the area shown in Figure 44,
i.e. half way between the two coarse blocks.
i=3

i=7

j=1

j=5

Kyte and Berry approximate the effective permeability using the arithmetic average
in each column, and then taking the harmonic average of the columns. The area
between the two coarse blocks is used, for reasons explained below.

28

Figure 44
The area used for
calculating the effective
absolute permeability

Permeability Upscaling

ki =

z k
j =1

ij

(20)

where zzj and Z are the thicknesses of the fine and coarse blocks, respectively. (In
this case, all the blocks are of equal size.)

kI =

x
i=3

ki

(21)

where xi and X are the lengths of the fine and coarse blocks, and k I is the required
effective absolute permeability.
The pseudos are then calculated, at certain times during the simulation. (These are
the times at which the restart files are written in the Eclipse simulation.)
2. Calculate the average water saturation:
5

Sw =

S
j =1 i =1
5
5

x i z j

w , ij ij

x z
j =1 i =1

ij

(22)

where ij is the porosity.


3. Calculate the total flow of oil and water out of the left coarse block (Figure 45).
5

q f = q f 5, j ,
j =1

(23)

where qf5,j is the flow of fluid f from fine block number (5,j).
i=5
j=1

Figure 45
Calculation of the total flow

j=5

4. Calculate the average phase pressures in the central column of each coarse
block. In this example, we use the fine blocks in columns 3 and 8, the shaded
areas in Figure 46.

Institute of Petroleum Engineering, Heriot-Watt University

29

i=3

i=7

j=1

j=5

Figure 46
The cells used for averaging
the phase pressures

II

In the Kyte and Berry method, the pressures are weighted by the phase permeabilities
times the height of the cells (which in this case are all the same size). This is so that
more weight is given to regions where there is greater flow. However, there is no
scientific justification for using this weighting. In the first coarse block (numbered,
I), the average pressure is:
5

P fI =

k
j =1

3j

k rf 3 z 3 j Pf 3 j gf (D3 j D)
5

k
j =1

3j

k rf 3 z 3 j

(24)

where D3j is the depth of cell (3,j) and D is the average depth of coarse cell I. The
term gf(D3j - D ) is to normalise the pressure to the grid block centre. The average
pressure for coarse block II is calculated in the same manner, but using column 8
instead of column 3. The pressure difference is then calculated as:

P f = P fI P fII .

(25)

5. The pseudo rel perms are then calculated using Darcys law. Firstly, calculate
the pseudo potential difference. (Potential is defined as = P-gz, so that the
flow rate is proportional to .)

f = P f gf
D,

(26)

where D is the depth difference between the two coarse grid centres. Then:

k rf =

f q f X
Zk I I

(27)

6. Calculate the pseudo capillary pressure using:

P c = P oI P wI

(28)

The Eclipse PSEUDO package can be used for calculating Kyte and Berry
pseudos.

30

Permeability Upscaling

2.6.3 Discussion on Numerical Dispersion

One advantage of pseudo-isation methods, such as that of Kyte and Berry is that they
can take account of numerical dispersion. When a simulation is carried out using a
larger grid, the front between the oil and water becomes more spread out. However,
the Kyte and Berry method counteracts this effect by calculating the flows on the
down-stream side of the coarse block, instead of the middle. This is illustrated by a
simple example. Figure 47 shows an example of input relative permeability curves
(rock curves).
0.9
0.8
0.7

Perm Rel

0.6
0.5
0.4
0.3
0.2
0.1
0

Figure 47
Example of rock curves

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Water Saturation

If the water saturation is Sw = 0.5, the rock curves show that there is a small amount

of oil and water flowing. However, when the average saturation, Sw , is 0.5 in the
coarse block, the distribution could be as shown in Figure 48.

oil
Figure 48
Example of the water
saturation in a coarse block

water

coarse
block

Since the water has reached only half way across the coarse block, there should be no
water flowing out of the right side. The Kyte and Berry method calculates the pseudo
relative permeabilities using the flow on the downstream side of the coarse block, to
prevent water breaking through too soon. The pseudo water relative permeability
curve is moved to the right, relative to the rock curves, as shown in Figure 49.

Institute of Petroleum Engineering, Heriot-Watt University

31

pseudos - solid lines


ave. rock curves - dashed lines

Relative Permeability

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Water Saturation

2.6.4 Disadvantages of the Kyte and Berry Method

There are certain problems with the Kyte and Berry method.

Negative rel perms are produced, if f has the same sign as q f .


Infinite rel perms occur if f is zero.
The method of averaging the pressures, using relative permeability as a weighting
function, may cause errors when the fluids are separated due to gravity. For
floods which are gravity-dominated, the TW method works better (Section
2.6.5).
Non-zero pseudo capillary pressure may be produced, even if there is no capillary
pressure in the fine-scale simulation. This is because a different weighting is
used for calculating the average pressure in each phase.
The capillary pressure may be different in different directions, because only the
central column is used for averaging the pressures.

Because of the first two disadvantages, i.e. negative, or infinite rel perms, pseudos
obtained from packages like the PSEUDO must be vetted before using at the coarse
scale. Often odd values of relative permeability are set to zero.
Good reviews of various methods for calculating pseudos are presented in Barker
and Dupouy (1999) and Barker and Thibeau (1997).
Note that dynamic upscaling methods, such as that of Kyte and Berry are difficult
to apply in practice. Ideally, a fine-scale two-phase flow simulation is required for
each coarse-scale cell (plus a flow jacket), and this is time consuming. Also, it is
difficult to determine the correct boundary conditions to use, so the results may not
be accurate.
If a pseudo is calculated for each coarse cell, in each direction, there may be 10,000s
of pseudos in the coarse-scale model. The number of pseudos must be reduced, by
grouping similar pseudos together.

32

Figure 49
Example of pseudo relative
permeability curves

Permeability Upscaling

Pseudo relative permeability curves depend on a number of factors, including:


(a) The balance of forces
The shape and end points of a pseudo depend on the ratio of viscous/capillary and
viscous/gravity forces. These ratios may be different in different parts of the reservoir.
(b) The well locations
The wells determine the flow rate and direction. If a new well is drilled, the pseudos
ought to be re-calculated.
Because of these problems, two-phase upscaling is rarely used for upscaling from a
geological model to a full-field simulation.
2.6.5 Alternative Methods

There are a number of similar methods to the Kyte and Berry Method.
(a) The Pore-Volume Weighted Method
The problems of non-zero capillary pressure and directional capillary pressure,
mentioned in Section 2.6.4, may be overcome by using a pore volume weighted
average of the pressures over the entire coarse block. ECLIPSE uses this method
for calculating the average capillary pressure in the Kyte and Berry method. Also,
pore volume weighting may be used for averaging the pressures when calculating
the pseudo relative permeabilities. In this case, the method is called the Pore Volume
Weighted Method. It is available in the Eclipse PSEUDO package.
(b) The TW Method
This method was developed by Nasir Darman at Heriot-Watt University (Darman et al,
1999). It is similar to the Kyte and Berry method, except transmissibility weighting is
used when calculating the average pressure. The method works better than the Kyte
and Berry method in cases where gravity effects are significant (e.g. a gas flood).
Both these methods share the same problems discussed in Section 2.6.4, namely,
they are difficult to apply in practice.
2.6.6 Example of the PVW Method

In two-phase dynamic upscaling methods, pseudo relative permeabilities are calculated, so that (hopefully) the results of a coarse-scale simulation provide a good
approximation to the fine-scale results. Layer 59 of the SPE 10 upscaling study
(Figure 37b) is used here as an example. A global simulation was performed on the
fine grid, i.e. the whole of the fine grid was included in the fine-scale flow simulation.
(Note that this would not be done, in practice, because there is no point in upscaling,
if you can simulation the whole fine grid.) The fine-scale model had 60 x 220 cells
and the coarse-scale model had 10 x 22, which corresponded to a scale-up factor of
6 x 10. Pseudo relative permeabilities were calculated for each coarse cell, in each
direction.
Figure 50 shows the oil saturation for the fine-scale simulation, a coarse-scale simulation
using the WDU method (Section 2.3.2), and a coarse-scale simulation using pseudos

Institute of Petroleum Engineering, Heriot-Watt University

33

from the PVW method. Both the WDU and the PWV methods give reasonable oil
saturation distributions. The oil recovery rate, for well P4, for these models is shown
in Figure 51, along with the oil rate for a coarse-scale simulation using single-phase
upscaling with local boundary conditions (Sections 2.2. and 2.3).
fine

WDU

PVW

Figure 50
The oil saturation
distribution for the finescale model of layer 59 and
the coarse-scale models
obtained using the WDU
and the PVW methods

Soil

0.20

0.35

0.50

0.65

0.80

Well Rate (m3/day)

60
50
40

Fine
Local
WDU
PVW

30
20
10
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Pore Volumes Injected

It can be seen in Figure 51 that both the WDU and the PVW methods agree well with
the fine-scale simulation. However, the results from the single-phase upscaling with
local boundary conditions are poor. This example shows that two-phase upscaling
is not necessarily always more accurate than single-phase upscaling.
2.7 Summary of Two-Phase Flow

The main points on two-phase flow are:

34

Two-phase upscaling is time consuming and not always robust, so is rarely used
by engineers.
Usually, only single-phase upscaling is performed.
But, heterogeneity interacts with two-phase flow, and tends to produce dispersion of
the flood front, which is not taken into account using single-phase upscaling.
So, single-phase upscaling may give rise to errors, especially when there is
a large scale-up factor, and the reservoir model is very heterogeneous (large
standard deviation).

Figure 51
The oil recovery rate for
well P4 for the fine-scale
model and coarse-scale
models obtained using
the WDU, PVW and local
upscaling methods

Permeability Upscaling

Errors in single-phase upscaling may be reduced by using non-uniform upscaling,


or well-drive upscaling.
Ideally, two-phase upscaling should be performed to take account of two-phase
flow.
Steady-state upscaling is relatively quick to apply, and is feasible for large
models. However, it is only valid in limited cases, e.g. when the fluids are
approximately in capillary equilibrium.
Dynamic methods are potentially more accurate.
The Kyte and Berry (1975) Method was described as an example.
Dynamic methods can compensate for the effects of numerical dispersion.
Dynamic methods are difficult to apply in practice.

3 ADDITIONAL TOPICS
This course has, so far, focussed mainly on common methods for upscaling a geological model for full-field simulation. Most of the single-phase upscaling methods
presented may be found in geological packages, such as IRAP/RMS and Petrel. (The
WDU and TW methods which were developed at Heriot-Watt are not available in
commercial packages.) However, there are a number of other important issues which
should be taken into account when upscaling. In this section, we cover these issues
in a variety of additional topics:
Upscaling as Wells
Permeability Tensors
The Geopseudo Method
Uncertainty and Upscaling
3.1 Upscaling at Wells

In the single-phase upscaling methods described in Chapter 1, we assumed that the


flow was linear. This means that the upscaling methods were not appropriate for
regions containing wells, where there is radial flow. We start with a brief overview
of simulation in blocks containing a well.
A grid block in the simulator is much larger than the diameter of a well, and the
pressure calculated for a block containing a well is different from the actual bottom
hole pressure. These are related by:

q=

Iw
(Pw Pb )

(29)

where Pw is the well-bore pressure and Pb is the pressure of the block. Iw is the well
index, given by:

Iw =

2kz
ln( ro rw )

Institute of Petroleum Engineering, Heriot-Watt University

(30)

35

where rw is the well-bore radius and ro is the equivalent radius, given by Peacemans
equation (Peaceman, 1978; Peaceman, 1983). Iw is also referred to as the well connection factor, or the connection transmissibility factor.
Durlofsky et al. (2000) put forward a method for upscaling in the near-well region.
Others have put forward similar methods. The method is only approximate, but
improves the accuracy of coarse-scale simulations. The first step is to calculate
effective single-phase permeabilities, using one of the conventional methods (e.g.
periodic boundary condition applied to each block in turn). Then, a fine-scale singlephase simulation of the well block and surrounding blocks is carried out (Figure 52).
From the results, the total flows out of the coarse-scale well block, and the average
pressures in the coarse blocks are calculated. These are used to calculate upscaled
transmissibilities between the coarse-scale well block and the surrounding blocks,
and a coarse-scale well index.
q

T4

T3

T1

T2
Figure 52
Near-well upscaling (after
Durlofsky et al., 2000)

This method improves the accuracy of upscaling at well, and it is also incorporated
into the well drive upscaling method (WDU), described in Section 2.3.2.
3.2 Permeability Tensors

Suppose that we have layers which are tilted at an angle to the horizontal, as in
Figure 53.
net flow in z-dir

x
net flow in x-dir

A pressure gradient has been applied in the x-direction. This will obviously give
rise to a flow in the x-direction. The fluid takes a path through the medium, so that
it expends a minimum amount of energy. There will be a component of flow up the
high permeability, and only a small amount of flow across the low permeable layer,
as shown. This gives rise to a net flow in the z-direction, or cross-flow. Here, the
term cross-flow is used to describe flow perpendicular to the applied pressure gradient. When calculating the effective permeability of this model, we need to take this
cross-flow into account. This may be done using a tensor effective permeability, k ,
where:
36

Figure 53
Cross-flow due to tilted
layers. Light-coloured
layers represent high
permeability and darkcoloured layers represent
low permeability

Permeability Upscaling

k xx

k = k yx
k zx

k xz

k yz
k zz

k xy
k yy
k zy

(31)

The first index applies to the flow direction, and the second to the direction of the
pressure gradient. For example kxy is the flow in the x-direction caused by a pressure gradient in the y-direction. The terms kxx, kyy, kzz are known as the diagonal
terms. These are the terms which are usually considered the horizontal and vertical
permeabilities, kh and kv, respectively. The other terms, which describe the crossflow, are the off-diagonal terms.
With tensor permeabilities, Darcys Law becomes:

u=

k
P,

(32)

where u is the Darcy velocity (vector) and P is pressure (scalar).

1
P
P
P
u x = k xx
+ k xy
+ k xz

x
y
z
1
P
P
P
u y = k yx
+ k yy
+ k yz

x
y
z
1
P
P
P
u z = k zx
+ k zy
+ k zz

x
y
z

(33)

In Sections 1.3.1 and 1.3.2, we studied flow along and across horizontal layers. The
model in Figure 5 is repeated here (Figure 54), showing the arithmetic average for
the effective permeability for along-layer flow and the harmonic average for across
layer flow.
x

Figure 54
The simple two-layer model

finefi
ne-sca
nescale
sca
le

coar
co
arse
ar
se--sca
se
scale
le

t1 = 3 mm, k1 = 10 m
mD
D
t2 = 5 mm
mm,, k 2 = 100 mD
mD

66.2
66
.25
.2
5 mD
22..86 mD
22
mD

In this model, there is no cross-flow, so we may write the effective permeability in


tensor form with zero off-diagonal terms, as follows:

0
66.25
k=
222.886
0
or in 3D:

Institute of Petroleum Engineering, Heriot-Watt University

37

0
0
66.25

k=
0
66.25
0

0
0
22.86
3.2.1 Flow Through Tilted Layers

x
z

z'

Figure 55
Layers tilted at an angle of
to the horizontal

x'

This model is essentially the same as the one in Figures 4 and 6, although the layers
have are repeated, and they have been tilted. In the frame of reference defined by the
x and z axes, the effective permeability may be calculated using the arithmetic and
harmonic averages as before. However, in the x-z co-ordinate system, the effective
permeability should be represented by a full tensor. The terms of the tensor may be
calculated from the arithmetic and harmonic averages, as follows:

k a cos2 + k h sin 2 ( k a k h ) sin cos


k=
2
2
( k a k h ) sin cos k a sin + k h cos

(34)

This formula is obtained by rotating the co-ordinate axes through an angle . (You
are not required to know the proof.) This example is in 2D, so only the kxx, kxz, kzx
and kzz are shown. Further rotations may be carried out around the x or z axes to
obtain a full 3D tensor.
Note that:
The tensor is symmetric (kxz = kzx).
Depending on the sign of , the off-diagonal terms may be positive or negative.
Example 5
Suppose the example in Figure 54 is rotated by 30 (Figure 56), and calculate the
effective permeability tensor.

38

Permeability Upscaling

'z
x

Figure 56
Layered model tilted by 30

cos230 = 0.75,

30o

sin230 = 0.25,

sin30.cos30 = 0.433.

From before, ka = 66.25 mD, and kh = 22.86 mD.

k xx = 66.25 0.75 + 22.86 0.25 = 55.40 mD,


k xz = (66.25 22.86) 0.433 = 18.79 mD,
k zz = 66.25 0.25 + 22.86 0.75 = 33.71 mD.
55.40 18.79
k=
.
18.79 33.71
Full tensor permeabilities may also be calculated from numerical simulations. It is
useful to use periodic boundaries, as described in Section 1.4.2. When a pressure
gradient is applied in the x-direction, there will be flow in the x-direction, and also
flow in the z-direction due to internal heterogeneity. These flows can be used to
calculate the kxx and kzx tensor terms. Then a pressure gradient is applied in the zdirection to obtain the kzz and kxz terms. (In 3D, a pressure gradient should also be
applied in the y-direction.)
3.2.2 Simulation with Full Permeability Tensors

Having calculated full effective permeability tensors, we need special software to


handle them at the larger scale. Conventional finite difference simulators use a 5-point
scheme in 2D and a 7-point scheme in 3D, and only take diagonal tensors e.g. when
running ECLIPSE, you usually specify PERMX, PERMY and PERMZ. Simulation
with full tensors is more complicated and more time-consuming, but some packages
allow the user to input full tensors. In Eclipse, there is a full tensor option which
allows you to specify terms such as PERMXY.
In 2D, a 9-point scheme is required to take account of cross-flow. This means that
there are 9 terms in each of the pressure equations, as illustrated in Equation (35).

a1Pi , j a 2 Pi 1, j a 3 Pi +1, j a 4 Pi , j 1 a 5 Pi , j +1
a 6 Pi 1, j 1 a 7 Pi 1, j +1 a 8 Pi +1, j 1 a 9 Pi +1, j +1 = 0.

(35)

The coefficients, ai, in Equation (35) depend on the transmissibilities between the
blocks. There are several different methods of discretisation which give slightly difInstitute of Petroleum Engineering, Heriot-Watt University

39

ferent results. To extend this to 3D, we need either a 19-point scheme or a 27-point
scheme. See Figure 57. (The 19-point scheme leaves out the 8 corners of the cube.)
Obviously, it takes longer to solve equations with a larger number of terms.
a)

b)

x
i-1,j-1

i,j-1

i+1,j-1

i-1,j

i,j

i+1,j

i-1,j+1

i,j+1

i+1,j+1

Figure 57
a) 9-point scheme for 2D.
b) 27-point scheme for 3D

Often the off-diagonal elements of the permeability tensor (kxy, etc) are negligible,
so the limitations of using a 5-point (2D) or a 7-point (3D) scheme are not serious.
In layered systems, the size of the off-diagonal term may be gauged from Equation
(34) in Section 3.2.1:

k xz = ( k a k h ) sin cos .

(36)

This is a maximum for = 45, and increases as (ka kh) increases. Therefore, full
permeability tensors become more important as the angle of the lamination or bedding increases, and as the permeability contrast increases.
3.3 Small-Scale Heterogeneity

Most reservoirs are modelled using, what is commonly termed a fine-scale geological
model. This is a stochastic model with grid cells of size approximately 50 m in the
horizontal directions, and about 0.5 m in the vertical. There are typically about 107
such cells in a full field model. These cells must be reduced in number to about 104 for
full-field simulation. However, each of the grid cells in the geological model is likely
to be heterogeneous, containing, for example, sedimentary structures. Petrophysical
data (permeabilities, relative permeabilities, and capillary pressures) are acquired from
core plugs, which are only a few cm long. When small-scale structure is present,
petrophysical data should be upscaled before being applied to the grid blocks of the
geological model. Figure 58 shows the ranges of scales of sedimentary structures,
along with the scales of measurements and typical sizes of models.

Vertical thickness (m)

100

Log

Flow model
Geological model

Core

0.1

0.01

Parasequences

Seismic data

10
0

Beds
Probe
Laminae

0.001
0.001

0.01

0.1

1.0

10

Horizontal length (m)

40

100

1000

10000

Figure 58
Length scales

Permeability Upscaling

For convenience, we consider upscaling as two separate stages (Figure 59). Stage
1 is upscaling from the smallest scale at which we may treat the rock as a porous
medium (rather than a network of pores), up to the scale of the stochastic geological model, i.e. from the mm cm scale to the m Dm scale. Stage 2 is upscaling
from the stochastic geological model to the full-field simulation model, which has
already been described.
Core plug

Sedimentary
Structure

Stage 1
Upscaling
Geological Model
~ 107 blocks

Figure 59
Two separate stages of
upscaling. (Geological
model taken from Tenth
SPE Comparative Solution
Project: A Comparison of
Techniques, by Christie
and Blunt, 2001.)

Stage 2
Upscaling

Simulation Model
~ 104 blocks

3.3.1 The Geopseudo Method

Upscaling from the core-scale to the scale of the geological model (Stage 1 in Figure
59) is frequently ignored by engineers, who apply core plug permeabilities and rock
relative permeability curves directly to the geological model. However, work carried
out at Heriot-Watt University has demonstrated that small-scale structures, such as
sedimentary lamination may have a significant effect on oil recovery (Corbett et al.,
1992; Ringrose et al., 1993; Huang et al., 1995). For example, in a waterflood of
a water-wet rock, water is imbibed into the low permeability laminae, and oil may
become trapped in the high permeability laminae.
The Geopseudo Method is an approach, where upscaling is carried out in stages,
using geologically significant length-scales (Figure 60). Models of typical sedimentary structures are created and permeability values are assigned to the laminae
(from probe permeameter measurements, or by analysing core plug data). Relative
permeabilities and capillary pressure curves are also assigned to each lamina-type (by
history matching SCAL experiments on core plugs). Flow simulations are carried
out to calculate the effective single-phase permeability and the two-phase pseudo
parameters. Additional stages of modelling and upscaling may be required e.g.
upscaling from beds to bed-sets.
In the finest-scale model, the grid cells may be a mm cube, or less. If we upscale to
blocks of 50 m x 50 m x 0.5 m, we are upscaling by a factor of at least 5 x 104 in the
horizontal directions and 500 in the vertical.

Institute of Petroleum Engineering, Heriot-Watt University

41

Low Perm

High Perm
Individual Rel. Perm Curves

Effective Perm

Pseudo Rel. Perm Curves

Figure 60
Illustration of the
Geopseudo Method

3.3.2 Capillary-Dominated Flow

At small scales the flow is often capillary-dominated. Figure 24 showed a moderate


capillary effect: the imbibition of water along the low permeability layer made the
flood front approximately level in the two layers (instead being ahead in the high
permeability layer, in the case of a viscous-dominated flood). In that model (Figure
21), the layers were 1 m thick, and the grid cells were 10 cm square. If the size of
the model is reduced by a factor 100, so that the layers are 1 cm thick, and represent
sedimentary laminae, the effects of capillary pressure are much stronger, as shown in
Figure 61. In this case, strong capillary imbibition draws water into the low permeability layers (black) so that the front advances faster in this layer. Notice that, in this
figure, there is little lateral variation in the shading, showing that the water saturation
is almost constant in each layer. This is because the front has been spread out by the
effects of capillary pressure, and the model is almost in capillary equilibrium.

Oil Saturation
0.3

0.4

0.5

0.6

0.7

When the flow is across the layers, as in Figure 62, the effects of capillary pressure
are even more striking. This figure shows the same small-scale model of sedimentary lamination. When the injection rate is low (average frontal advance rate of
0.3 m/day), the flood is capillary-dominated (Figure 62a), water (black) has been
imbibed into the low permeability layers leaving oil trapped in the high permeability
laminae (grey). As the injection rate is increased, the oil has more viscous force and
can overcome the capillary forces leading to less trapping of oil (Figures 61b). In
the case of Figure 62c, the flood is viscous dominated and all the movable oil has
been displaced by water.

42

Figure 61
Example of capillarydominated flood in a
layered model

Permeability Upscaling

b)

a)

Figure 62
Examples of acrosslayer flow. a) capillary
dominated, b) intermediate,
c) viscous-dominated

c)

Oil Saturation
0.3

0.4

0.5

0.6

0.7

The examples shown in Figures 61 and 62 demonstrate the significance of capillary


effects at the small-scale. When upscaling from the lamina-scale, these effects should
not be ignored, and two-phase upscaling should be performed.
3.3.3 Geopseudo Example

All the upscaling methods described in the previous sections may be used in the
Geopseudo approach, depending on the type of heterogeneities and the fluids flowing
averaging, single-phase numerical methods, two-phase steady-state methods, or
two-phase dynamic methods. Since a flood is often capillary-dominated, as shown
above, steady-state upscaling using the capillary equilibrium method is often appropriate. Two-phase dynamic upscaling may also be used, and we show an example
of the Kyte and Berry method below.
Figure 63 shows a model of sedimentary ripples. Kyte and Berry pseudos were calculated for the model using four different flow rates. There is a factor of 10 between
each flow rate, with rate 1 being the fastest. Figure 64 shows the resulting pseudos
(from Pickup and Stephen, 2000).
Figure 63
A model of ripples (based
on the Ardross Outcrop,
near St. Monance in Fife,
Scotland)

10 mD

3 cm, 54 cells

200 mD

1 cm,
18 cells

0.9

0.9
0.7
0.6

Relative Permeability

Relative Permeability

Figure 64
Pseudo relative
permeabilities for different
flow rates, for oil (left) and
water (right)

0.8

rate 1
rate 2
rate 3
rate 4

0.8

0.5
0.4
0.3
0.2
0.1
0.0
0.2

0.3

0.4

0.5

0.6

0.7

0.8

rate 1
rate 2
rate 3
rate 4

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.2

Water Saturation

Institute of Petroleum Engineering, Heriot-Watt University

0.3

0.4

0.5

0.6

0.7

0.8

Water Saturation

43

Note the following:


1. At high rates, the pseudos are shifted to the right. This is to compensate for
numerical dispersion.
2. At very low flow rates (rate 4), the flood is capillary-dominated, and the oil is
trapped. The pseudo oil relative permeability goes to zero around Sw = 0.46.
3.3.4 When to use the Geopseudo Method

Geopseudo upscaling may be time-consuming, and there is no point in upscaling from


the smallest scales, unless cores are available for the field. Cores must be studied to
identify the sedimentary structures present, and probe permeability measurements
should be taken to populate the small-scale models. Additionally, reliable SCAL
data is also required.
Ringrose et al. (1999) give a list of guidelines for when Geopseudo upscaling may
be necessary:
1) Are immiscible fluids flowing?
2) Are significant small-scale heterogeneities present? Specifically:
Is the permeability contrast greater than 5:1?
Is the layer thickness less than 20 cm?
Is the mean permeability less than 500 mD?
3) What is the large-scale structure of the reservoir? In many cases, large-scale
connectivity may be the dominant issue, in which case, small-scale structure
may have to be ignored. The Weber and van Geuns (1990) classification may
be used to describe the large-scale structures:
Layer cake reservoirs small-scale structure will usually have primary
importance.
Jigsaw puzzle reservoirs small-scale structure may be important.
Labyrinth reservoirs small-scale structure will usually be of secondary
importance.
3.4 Uncertainty and Upscaling

During the 1990s, reservoir modelling developed (along with computing power) so
that geologists could create models containing millions of grid cells. Such models are
often time-consuming to generate, and only a few are created for each reservoir. These
detailed models are too large for full-field flow simulation, and must be upscaled to
reduce the number of cells to, about 104 or 105. Research into upscaling has focussed
on trying to develop methods to accurately upscale these types of models.
However, it is now recognised that there are many uncertainties in the reservoir
modelling, and instead of concentrating on a few detailed models, geologists and
engineers are starting to generate thousands of models in order to characterise the
effects of uncertainty. These models must be coarse so that the simulations can run
very quickly.
These changes mean that, in future, people are less likely to follow the traditional
upscaling approach. However, if the effects of fine-scale structure are ignored, this

44

Permeability Upscaling

will lead to errors in the predicted recovery. It is therefore very important to understand
the effects of possible sub-grid heterogeneity on absolute and relative permeability,
and to include these effects, when necessary. This is an area of active research at
Heriot-Watt University.
3.5 Upscaling Summary

Several reviews have been published on upscaling. These give an overview of some
of the methods described in this chapter: e.g. Christie (1996), Renard and Marsily
(1997) and Christie (2001).
Here is a summary of the main points:

The effective permeability of simple permeability models (layered or random)


may be calculated using averaging.

In general, effective permeability should be calculated using a numerical


simulation, along with suitable boundary conditions.

Permeability upscaling is often inaccurate, particularly when the coarse cell size
is comparable, or slightly larger than the correlation length of the permeability
distribution, and when the standard deviation is large.

Usually only single-phase upscaling is used in two-phase systems. However,


this can give rise to errors, especially when the scale-up factor is large and when
the standard deviation of the permeability distribution is large.

Single-phase upscaling for two-phase systems may be made more accurate by


using a non-uniform coarse grid, or by using the Well Drive Upscaling method,
which increases the accuracy of single-phase upscaling by using the correct
boundary conditions.

The capillary equilibrium method is useful, particularly for small-scale models.


It is feasible, even for models with a relatively large number of grid cells.

Two-phase dynamic upscaling methods should be able to reproduce two-phase


flow on a coarse scale. The Kyte and Berry (1975) method is an example of a
pressure averaging approach.

In general, two-phase upscaling is difficult to apply. It is more time consuming


than single-phase upscaling, and the results are not robust (negative or infinite
values may be obtained).

The regions around wells should be treated as a special case, because the flow
is radial. The well index and the transmissibilities around the well block should
eb upscaled.

Permeability is actually a tensor quantity (4 terms in 2D, 9 terms in 3D). Full


tensors may be used to take account of cross-flow within a grid cell. However,
in general, only the diagonal terms are used (kxx, kyy and kzz, often referred to
as kx, ky and kz).

Institute of Petroleum Engineering, Heriot-Watt University

45

It is important to take account of small-scale (mm m) heterogeneity in some


reservoirs. This may be done using the Geopseudo Method, in which models
of sedimentary structures are generated and upscaled.

Capillary effects are often significant at small-scales, and it is important to take


these into account using two-phase upscaling (steady-state or dynamic).

46

Permeability Upscaling

4 REFERENCES
Barker, J. W. and Thibeau, S., 1997. A Critical Review of the Use of Pseudo Relative
Permeabilities for Upscaling, SPE Reservoir Engineering, May, 1997, 138-143.
Barker, J. W. and Dupouy, P., 1999. An Analysis of Dynamic Pseudo-Relative Permeability Methods for Oil-Water Flows, Petroleum Geoscience, 5 (4), 385 - 394.
Christie, M. A., 1996. Upscaling for Reservoir Simulation, J. Pet. Tech., November
1996, 48, 1004-1008.
Christie, M. A., 2001. Flow in Porous Media Scale Up of Multiphase Flow,
Current Opinion in Colloid and Interface Science, 6, 23 241.
Christie, M. A. and Blunt, M. J., 2001. Tenth SPE Comparative Solution Project: A
Comparison of Upscaling Techniques, presented at the SPE Reservoir Simulation
Symposium, Houston Texas, 11 14 February, 2001.
Corbett, P. W. M., Ringrose, P. S., Jensen, J. L. and Sorbie, K. S., 1992. Laminated
Clastic Reservoirs - The Interplay of Capillary Pressure and Sedimentary Architecture, SPE 24699, presented at the 67th Annual Technical Conference of the SPE,
Washington, DC, 4 - 7 October, 1992.
Darman, N. H., 2000. Upscaling of Two-Phase Flow in Oil-Gas Systems, Ph.D.
Thesis, Heriot-Watt University.
Darman, N. H., Pickup, G. E. and Sorbie, K. S., 2002. A Comparison of Two-Phase
Dynamic Upscaling Methods Based on Fluid Potentials, Computational Geosciences,
6, 5 27.
Durlofsky, L. J., Behrens, R. A., Jones, R. C. and Bernath, A., 1996. Scale Up of
Heterogeneous Three Dimensional Reservoir Descriptions, SPEJ, 1, 313-326.
Durlofsky, L. J., Jones, R. C. and Milliken, W. J., 1997. A Nonuniform Coarsening
Approach for the Scale Up of Displacement Processes in Heterogeneous Porous
Media, Advances in Water Resources, 20, 335 347.
Durlofsky, L. J., Milliken, W. J. and Bernath, A., 2000. Scaleup in the Near-Well
Region, SPEJ, 5 (1), 110 117.
Huang, Y., Ringrose, P. R. and Sorbie, K. S., 1995. Capillary Trapping Mechanisms
in Water-Wet Laminated Rocks, SPE RE, 10 (4), 287 292.
Kyte, J. R. and Berry, D. W., 1975. New Pseudo Functions to Control Numerical
Dispersion, SPEJ, August 1975, 269-276.
Peaceman, D. W., 1978. Interpretation of Well-Block Pressures in Numerical Reservoir Simulation, SPEJ, June 1978, 183-194.

Institute of Petroleum Engineering, Heriot-Watt University

47

Peaceman, D. W., 1983. Interpretation of Well-Block Pressures in Numerical


Reservoir Simulation with Nonsquare Grid Blocks and Anisotropic Permeability,
SPEJ, June 1983, 531-543.
Pickup, G. E. and Stephen, K. D., 2000. An Assessment of Steady-State Scale-Up
for Small-Scale Geological Models, Petroleum Geoscience, 6 (3), 203 210.
Pickup, G. E., Ringrose, P. S. and Sharif, A., 2000.Steady-State Upscaling: From
Lamina-Scale to Full-Field Model, SPEJ, 5 (2), 208 217.
Renard, P. and de Marsily, G., Calculating Equivalent Permeability: A Review,
Advances in Water Resources, 20 (5/6), 253 278.
Ringrose, P. S., Sorbie, K. S., Corbett, P. W. M. and Jensen, J. L., 1993. Immiscible
Flow Behaviour in Laminated and Cross-bedded Sandstones, J. Petroleum Science
and Engineering, 9(2), 103-124.
Ringrose, P. S., Pickup, G. E., Jensen, J. L. and Forrester, M. M., 1999. The Ardross
Reservoir Gridblock Analog: Sedimentology, Statistical Representivity, and Flow
Upscaling, in Reservoir Characterization Recent Advances, eds R. Schatzinger
and J. Jordan, AAPG Memoir 71, p 256 276.
Stone, H. L. 1991. Rigorous Black Oil Pseudo Functions, SPE 21207, presented
at the 11th SPE Symposium on Reservoir Simulation, Anaheim, CA, February, 1720, 1991.
Weber, K. J. and van Geuns, L. C., 1990. Framework for Constructing Clastic Reservoir Simulation models. JPT, October 1990, p 1248 1297.
Zhang, P., Pickup, G. E. and Christie, M. A., 2005. A New Upscaling Approach
for Highly Heterogeneous Reservoirs, presented at the SPE Reservoir Simulation
Symposium , February 2005.

48

Petrophysical Input

1 INTRODUCTION

7 CONCLUDING REMARKS

2 MODELLING SINGLE-PHASE FLOW AT THE


PORE-SCALE - A BRIEF OVERVIEW
2.1 Deviations from Darcys Law
2.2 Empirical Models
2.3 Probabilistic Models
2.4 Capillary Bundle Models
2.5 First Principles Derivation of CarmenKozeny Model
2.6 Network Modelling Techniques

8 APPENDIX A:
and Concepts

3 MODELLING
MULTIPHASE
FLOW
AT THE PORE-SCALE
3.1 Capillary Pressure What Does it Mean
and When is it Important?
3.2 Steady-and Unsteady-State Flow
3.3 Drainage at the Pore-Scale
3.4 Imbibition at the Pore-Scale
3.5 The Pore Doublet Model
3.6 Introduction to Percolation Theory
3.7 Network Modelling of Multiphase Flow
4 EXPERIMENTAL DETERMINATION OF
PETROPHYSICAL DATA
4.1 Laboratory Measurement of Capillary
Pressure
4.2 Laboratory Measurement of Relative
Permeability
5 EMPIRICAL AND THEORETICAL APPROACHES
TO GENERATING PETROPHYSICAL
PROPERTIES FOR RESERVOIR SIMULATION
5.1 Methods for Generating Capillary Pressure
Curves and Pore Size Distributions
5.2 Methods for Generating Relative
Permeabilities
5.3 Hysteresis Phenomena
6 WETTABILITY - CONCEPTS AND
APPLICATIONS
6.1 Introductory Concepts
6.2 Wettability Measurement and Classification
6.3 The Impact of Wettability on Petrophysical
Properties
6.4 Network Modelling of Wettability Effects

Some Useful Definitions

9 APPENDIX B: Unsteady-State Relative


Permeability Calculations
10APPENDIX C: Details of the Heriot-Watt MixWet
Simulator

LEARNING OBJECTIVES
Having worked through this chapter the students should be able to...
Modelling Single Phase Flow
Appreciate the different types of model used to predict single phase permeability
Use a Carman-Kozeny equation to calculate absolute permeability given values
for the remaining variables
Modelling Multiphase Flow
Explain the meaning of capillary pressure and use Laplaces equation to relate
capillary pressure to pore entry radius, contact angle and interfacial tension
Identify the difference between steady- and unsteady-state flow
Describe the pore-scale physics characterising drainage processes in porous
media
Describe the pore-scale physics characterising imbibition processes in porous
media
Describe a network model and explain how it can be used to investigate
multiphase flow in porous media
Experimental Determination of Petrophysical Data
Identify several different methods for measuring capillary pressure and relative
permeability
Generating Petrophysical Properties for Reservoir Simulation
Calculate oil-water capillary pressure curves from mercury injection data
Derive a pore size distribution from mercury injection data
Generate several capillary pressure curves from a single curve using a LeverettJ function
Determine the Brooks and Corey _-parameter from capillary pressure data and
use this to predict relative permeabilities given the relevant equations
Identify three causes of capillary pressure hysteresis
Wettability - Concepts and Applications
Explain how wettability variations affect waterflooding at the pore-scale
Identify two wettability measures used routinely in industry
List Craigs Rules of Thumb in the context of water-wet and oil-wet relative
permeability curves

Petrophysical Input

1 INTRODUCTION
Outline of the purpose of this chapter:
To inform the student of the types of petrophysical data that are used in reservoir
simulation (k, , kro, krw, Pc). k and are generally measured or determined by
correlation or model, so, the central focus here will be on multi-phase properties (kro,
krw, Pc);
To briefly review relative permeability and capillary pressure are measured
experimentally.
To explain the underlying pore-scale physics of two-phase flow and show how this
behaviour leads to the results we see at the macroscopic scale.
To review empirical and theoretical models used to generate relative permeabilities
and their application in Reservoir Simulation;
To review wettability measures (such as USBM and Amott tests) and to explain
how wettability modifies relative permeability and capillary pressure.
The notion of a "porous medium" immediately conjures up an intuitive picture: put
in its crudest terms, a porous medium may be thought of as a solid with holes in it.
Unfortunately, such a superficial definition is of little use when trying to describe such
materials objectively, and a more precise formulation must be attempted. A cylindrical
pipe, for example, would not generally be considered a porous medium, nor would a
solid containing isolated holes. There is a tacit understanding that "real" porous media
should be capable of sustaining fluid transport, implying a certain degree of
interconnectedness within the underlying pore structure. In short, a truly porous
material should have a specific permeability associated with it.
There are countless examples of porous materials in everyday life, each with its own
particular pore structure and transport potential. These range from leather, wood,
paper and textiles, to bricks, concrete and sand; even animal tissue and bones contain
intricate pore networks. The need to understand such a vast array of permeable
materials has consequently fostered a great deal of scientific interest from many
diverse fields: soil mechanics, groundwater hydrology, industrial filtration, and
petroleum engineering, to name but a few. Although the theme of fluid flow through
porous media is a common feature of all of the disciplines listed above, each has its
own technical terminology associated with the subject. For example, "dewetting",
"desaturation" and "drainage" are all terms synonymous with the displacement of a
wetting phase from the interstices of a porous material by a nonwetting phase.
Throughout this section, however, the terminology used will be that generally
encountered in the petroleum industry. A variety of fundamental concepts relating to
both pore structure and solid/fluid interactions can be found in Appendix A.

Institute of Petroleum Engineering, Heriot-Watt University

2 MODELLING SINGLE-PHASE FLOW AT THE PORE-SCALE


- A BRIEF OVERVIEW
Examination of any photomicrograph immediately demonstrates why the modelling
of fluid flow through a porous medium is such a formidable task: the underlying pore
structure is extremely complex, with tortuous channels embedded throughout the
solid matrix. Nevertheless, over the years there have been many attempts to
encapsulate this structure into a simple, idealised analogue. Such models can
generally be divided into four broad categories; (i) those which attempt to reduce the
porous medium to a single representative conduit, (ii) probabilistic models, (iii)
empirical correlations, and (iv) network models, where the medium is approximated
by a lattice of connected conduits with distributed radii. A brief discussion of such
analogues will be presented below; more detailed descriptions can be found in the
monographs of Scheidegger (1963) and Dullien (1979).
2.1 Deviations from Darcys Law

Although Darcys Law has been validated countless times experimentally, we should
nevertheless be aware of some possible difficulties. For example, liquid permeabilities
can be greatly affected by clay distribution, brine composition, and brine pH. This is
evident in Table 1, which shows variations in absolute permeability with increased
brine salinity in some cases, decreased salinity leads to a decreased permeability
estimate, whilst other samples exhibit the reverse behaviour. There is also experimental
evidence for so-called lubrication effects, where oil permeability measured at Swi
actually exceeds kabs (this is rather surprising when one considers the fact that isolated
water islands within a sample should actually form effective baffles to flow).

Field

Zone

Ka

K1000

K500

K300

K200

K100

KW

S
S
S
S
S

34
34
34
34
34

4080
24800
40100
39700
12000

1445
11800
23000
20400
5450

1380
10600
18600
17600
4550

1290
10000
15300
17300
4600

1190
9000
13800
17100
4510

885
7400
8200
14300
3280

17.2
147
270
1680
167

S
S
S
S
S

34
34
34
34
34

4850
22800
34800
27000
12500

1910
13600
23600
21000
4750

1430
6150
7800
15400
2800

925
4010
5460
13100
1680

736
3490
5220
12900
973

326
1970
3860
10900
157

5.0
19.5
9.9
1030
2.4

S
S
S
S
T

34
34
34
34
36

13600
7640
111000
6500
2630

5160
1788
4250
2380
2180

4640
1840
2520
2080
2140

4200
2010
1500
1585
2080

4150
2540
866
1230
2150

2790
2020
180
794
2010

197
119
6.2
4.1
1960

T
T
T
T
T

36
36
36
36
36

3340
2640
3360
4020
3090

2820
2040
2500
3180
2080

2730
1920
2400
2900
1900

2700
1860
2340
2860
1750

2690
1860
2340
2820
1630

2490
1860
2280
2650
1490

2460
1550
2060
2460
1040

Table 1
Effect of water salinity on
permeability of natural
cores (grains per gallon of
chloride ion as shown). Ka
means permeability to air;
K500 means permeability
to 500 grains per gal
chloride solution; Kw
means permeability to
fresh water

Petrophysical Input

There are two other well-known limitations of Darcys Law:


(i) At high injection rates, inertial effects can become important and Darcys Law
should be replaced with the Forcheimer Equation:

dp
= aq + bq 2
dx
where is the fluid viscosity and the second term on the right-hand side corresponds
to inertial effects.
(ii) Low pressure gas measurements must be corrected by using the Klinkenberg
Equation:


k app = k 1 +
p

where kapp is the apparent (measured) permeability, k the actual permeability, p the gas
pressure, and a parameter that depends upon the properties of the gas being used.
The correction is needed because the mean free path of a gas molecule at low pressure
is of the order of a pore radius and the continuum concept begins to break down.
2.2 Empirical Models

There have been numerous attempts to derive correlations between permeability and
other sample properties, such as porosity, capillary pressure, grain size distribution,
and electrical properties. The porosity/permeability relationship has perhaps been the
most widely studied (see, for example, Mavis and Wilsey, 1936; Bche, 1937; Rose,
1945; Habesch, 1990), but correlations vary so much that no universally accepted
formula can be adopted successfully. Consequently, most empirical correlations
contain "geometrical factors" which serve to fit experimental measurements without
any real consideration of the underlying pore structure.
2.3 Probabilistic Models

As their name suggests, probabilistic models involve the use of some kind of
probability law. One of the most popular of these is the "cut-and-random-rejoin"
model of Childs and Collis-George (1950), which has been further extended by
Marshall (1958) and Millington and Quirk (1961). The underlying theory involves the
sectioning of the porous sample into two parts perpendicular to the direction of flow.
These are then joined together again in a random fashion, and statistical analysis is
used to approximate the permeability of the subsequent hybrid.
2.4 Capillary Bundle Models

Capillary bundle models characterise the porous medium using systems of capillary
tubes with well-defined properties (Figure 1). Each different model contains tubes
with different characteristics: they may be uniform and identical, for example, or
uniform but with distributed radii, or periodically constricted and identical, etc.

Institute of Petroleum Engineering, Heriot-Watt University

k=

D2
32

Dmax

D2 =

f(s)ds

Dmin
2

k=

96

<>
D

k=

[ f(D) dD / D ]

96

f(D) dD / D6

+ 2D

If all tubes are identical (diameter=D) and lie parallel to the flow direction, then a
combination of Poiseuille's law and Darcy's law gives:

k=

D2
32

(1)

as a permeability predictor, where is the porosity (see Appendix B for mathematical


details).
An interesting aspect of this simple result is that the quantity (k/)1/2 can be thought
of as a sort of average pore diameter.
2.5 First Principles Derivation of Carmen-Kozeny Model

This popular approach, based upon hydraulic radius theory, relies upon two main
assumptions: (i) that a porous medium can be adequately characterised by a single
tortuous channel having a characteristic radius, usually called the hydraulic radius,
and (ii) that the effect of the interconnected pore structure can be contained within an
empirical constant known as the tortuosity factor. The mathematical details of this
approach are given in Appendix B.
Carmen-Kozeny Model - One of the more notable correlations based upon conduit
flow is that developed by Carmen and Kozeny (Carman, 1937, 1938, 1956; Kozeny,
1927). The basic premise of this modelling approach is that particle transit times in the
actual porous medium and the equivalent tortuous rough conduit must be the same.
After some analysis (see Appendix B), we arrive at the relationship:

k=

3
2 T 2 (1 )2 Ss2

(2)

See Appendix A for definitions


The permeability can be written in terms of an average grain diameter (Dp) by noting
that, for spherical particles, Ss=6/Dp. Hence,

Petrophysical Input

k=

3D 2p
72 T 2 (1 )2

(3)

Although a whole family of similar models exist, they differ only in the method of
calculating an hydraulic radius RH and shape factors.
2.6 Network Modelling Techniques

Models that account for the interconnected nature of porous media constitute a group
of analogues which can truly be referred to as network models. Here, the medium is
modelled using a system of interconnected capillary elements, which generally
configure to some known lattice topology. A variety of lattices are shown in Figure
2. Although these network structures are somewhat idealised, the capillary radii are
assigned randomly from a realistic pore size distribution in an effort to partially
reconstruct the actual porous medium under investigation.

Figure 2

Hexagonal

Square

Kagome

Triagonal

Cubic

Crossed square

A fully interconnected network was first used by Fatt (1956) for primary drainage
studies. Although the two-dimensional lattice used was extremely small (200-400
tubes), the novelty of the approach encouraged great interest in the subject, and
improvements on Fatt's primitive model were soon forthcoming (Rose ,1957; Dodd
& Kiel ,1959). However, simulations using small lattices could never hope to capture
the full behaviour of microscopic flow processes. With the advent of high speed
computers, however, much larger 3D systems can now be constructed, with the result
that microscopic flow behaviour can be more accurately simulated. Figure 3 shows
a capillary dominated drainage process using an 80 x 80 square lattice, the details of
which will be presented later: the resulting picture is somewhat more reassuring. The
fractal capillary fingering is in excellent agreement with experimental observation
(see Lenormand et al, 1988).

Institute of Petroleum Engineering, Heriot-Watt University

Figure 3

Only recently has it been computationally feasible to carry out studies using large
three-dimensional networks (Figure 4).

Figure 4

The Basic Model


Many network models attempt to distinguish between "pores" and "throats", by
building networks consisting of hollow spheres connected by thin capillary tubes. In
such models, all of the liquid volume is assumed to be contained in the spherical pores
with pressure differences being maintained by the throats (Lenormand, 1986). The
approach taken here is somewhat more straightforward. The porous medium is
initially modelled using a three-dimensional cubic network of what will be referred
to as pore elements; unlike many previous studies, no distinction is made here between
pores and throats. The model consists of a three dimensional cubic network of
8

Petrophysical Input

capillary elements. This simple lattice has dimensions Nx x Ny x Nz where Nx, Ny, Nz
are the number of nodes in the x, y and z directions respectively. Periodic boundary
conditions are assumed in the y and z directions in order to simulate larger systems and
eliminate surface effects. The pressure gradient is taken to be in the x direction.
Now, for a single element of radius r and length L, the flow Q is given by Poiseuilles
law:

Q=

r 4 P
8 L

(4)

where is the viscosity and P the pressure difference acting across the capillary. At
each node (i, j, k), the sum of the flows Qi must add up to zero (conservation of mass),
and so:
6

=0

i =1

Consideration of the whole network leads to a set of Nx.Ny.Nz linear pressure


equations, the solutions to which can then used to calculate the elemental flows.
Summing the outlet flows yields the total network flow, which can then be substituted
into Darcy's equation to give a value for the total network permeability. If the system
contains more than one fluid, then the process is carried out for each fluid in turn; the
resulting phase conductivities now being referred to as effective permeabilities. The
modelling of multiphase flow is discussed more fully in later sections.
3 MODELLING MULTIPHASE FLOW AT THE PORE-SCALE
3.1 Capillary Pressure - What Does it Mean and When is it Important?

For many, the term "capillary pressure" is a rather difficult concept to grasp, especially
in the context of a flowing hydrocarbon reservoir. For example, we could be shown
the schematic in Figure 5(a) and wonder why no oil flow occurs even though Poil>Pwater.
Athough capillary pressure may seem problematic, we shall soon see that it is actually
a very straightforward measure to interpret.

Institute of Petroleum Engineering, Heriot-Watt University

(a)

Poil>Pwater - why no flow?

Oil

Water

Projected
area
= r2
(b)

(c)

P0
Pi

(Pi - P0)R2 = 2R

r
=> (Pi - P0) =

2
R

Figure 5

Let us first consider a rubber balloon that has been inflated to a certain pressure (Pi)
and then tied (we will assume that this "experiment" is taking place in an atmosphere
at pressure Po, usually atmospheric pressure). If a force balance is considered for one
half of the balloon (Figure 5(c)), then we can show that, at equilibrium, the elastic
force acting around the circular perimeter of the balloon must counterbalance the
difference in pressure projected onto the shaded cross section.
This leads to a relationship between the pressure difference across the balloon surface
and the radius of the balloon:

( Pi Po ) =

2
R

(5)

where is the elastic tension characterising the balloon wall (dimensions of Force/
Length). Here is an example where a pressure difference exists between two regions
of fluid but no flow occurs the elastic membrane of the balloon counteracts this.
Now, instead of thinking of a rubber balloon (where is actually a function of R itself),
we can carry out a similar analysis for a gas bubble at pressure Pgas floating in oil at
pressure Poil. We can now write immediately:

( Pgas Poil ) =

2 go
R

(6)

where go represents the interfacial tension between gas and oil. This relationship tells
us that the pressure difference across a small spherical bubble is larger than that across
a large spherical bubble.
This type of analysis can be applied to the more general case of an interface
characterised by two different principal radii of curvature (eg a sausage-shaped
balloon Figure B2). The details are slightly more involved (see Dullien (1979), but
the final result is simply:
10

Petrophysical Input

( Pgas Poil ) = go (

1
1
+
)
R1 R 2

(7)

where R1 and R2 are the principal radii of curvature characterising the interface.
Return now to the idealized (and some what unrealistic) situation where we have two
fluids at equilibrium in a capillary tube separated by a curved interface (this curved
interface appears at the microscopic scale when one of the fluids preferentially wets
a solid surface, Figure 6). The pressure difference across the fluid-fluid interface is
known as the capillary pressure. Note the difference between tube radius (rA) and the
interface radius (RA) when the contact angle is non-zero.
RA

rA

Figure 6

A little bit of trigonometry can be used to derive an equation for two fluids at
equilibrium in a circular cylinder, known as the Young-Laplace equation. This relates
the pressure difference across a curved interface (i.e. capillary pressure Pc) in terms
of the associated contact angle, interfacial tension and pore (tube) radius:

Pc = ( Pgas Poil ) = go cos (

2 go cos
1 1
+ )=
R R
R

(8)

Although we have used gas displacing oil in our example, results for any nonwetting
fluid displacing a wetting fluid can be inferred immediately.
Another consequence of the equation is that larger pressure differences are needed for
a nonwetting phase to displace a wetting phase from smaller tubes ((Pgas-Poil)1/R).
We can therefore go on to examine the very simple drainage process shown in Figure
7, where oil (dark) displaces water (light) from a set of parallel tubes (R1>R2>R3) as
oil pressure is gradually increased. Once the oil pressure exceeds the water pressure
by an amount 2owcos/R1, then the largest tube fills and the system settles down to
a new equilibrium configuration. Subsequent displacements occur at (PoilPwat)>2owcos/R2, and (Poil-Pwat)>2owcos/R3. The corresponding plot of water
saturation vs capillary pressure is shown in Figure 8. This could be thought of as a very
basic capillary pressure curve and we will return to this concept later.

Institute of Petroleum Engineering, Heriot-Watt University

11

P1

P2

P3

P4

Figure 7

Pc
2 cos
R3
2 cos
R2
2 cos
R1
Sw

Simulators use capillary pressure curves to relate oil and water pressures at a given
water saturation (see earlier chapters regarding this). Generally, capillary pressure is
mostly important at the small scale (~cm), although it should also be included in
reservoir-scale models involving transition zones.
3.2 Steady-and Unsteady-State Flow

The aim of this section is to broaden the understanding of two-phase flow in porous
media; more specifically, to the simultaneous flow of oil and water through reservoir
rock. To this end, concepts from percolation theory will be introduced towards the end
of this section, where the physical porous medium will be approximated using an
interconnected capillary network and each element may contain a different fluid:
either water or oil. Before dealing directly with percolation issues, however, it will be
beneficial to first discuss some of the more general aspects of two-phase flow, such
as phase distributions and relative permeabilities.
When dealing with any process involving multiphase flow, it is important to distinguish
between steady-state and unsteady-state regimes (Figure 9). With regard to flow in
porous media, the former is characterised by phase saturations that are invariant with
time; that is, the volume flux of each fluid entering the system is the same as that
leaving. Experimentally, this would be achieved by fixing the inlet flows of oil and
water at a certain ratio and leaving the system to equilibrate (such that the individual
fluid fluxes exiting the sample are the same as those entering). There are consequently
no pore-scale displacements of any kind once steady-state has been reached - each
12

Figure 8

Petrophysical Input

phase tends to flow within its own tortuous network of pores. Once measurements
have been completed at this ratio of fluxes, a new ratio could be set and the process
repeated.
Core-Scale

Steady-state

Unsteady-state
Pore-Scale

Figure 9

Steady-state

Unsteady-state

In contrast to this, unsteady-state flow is accompanied by almost continuous saturation


changes, implying continuous displacement of one fluid by another. This would occur
if one fluid was injected into a sample containing a second "resident" phase. Note that
the two flow regimes are seldom independent, however, as steady-state conditions are
only achieved once some degree of transient unsteady-state displacement has taken
place to redistribute the phases.
Although the "channel flow concept" of steady-state flow outlined above appears to
be generally valid (Craig, 1971), there are certain conditions under which this
assumption may be questioned. If there is no strong wetting preference for the matrix,
for example, or if the interfacial tension between the two fluids is very small, then a
slug-like flow regime may develop. Alternatively, if the flow channels are rough and
have an irregular cross-section (which is almost always the case in natural rock
material), then the wetting phase will tend to line the channel walls and the nonwetting
phase will usually reside in the centre of pores. Such phenomena can play a vital role
in determining phase distributions during a variety of displacements, and their effects
cannot always be overlooked. We shall return to these Issues later.
3.3 Drainage at the Pore-Scale

In order to better understand the following discussion on two phase displacements, we


begin with the idea of a pore size distribution, PSD a schematic representation of
the range and frequency of pore radii characterising a given sample (Cf probability
distributions from statistics). A schematic PSD and 3D network are shown in Figure
10, where the network of "pores" are simply capillary tubes of different radius (but
equal length in this case). The radius r of each pore is the capillary entry radius (i.e.
the radius characterising the pressure difference required for a nonwetting fluid to
invade a tube containing a wetting fluid, see Section 3.1).
We will now discuss drainage capillary pressure and relative permeability in two
different arrays of capillary tubes. First, we will return to a capillary bundle model and
then we will go on to examine drainage in an interconnected network.

Institute of Petroleum Engineering, Heriot-Watt University

13

(a)

(b)
PSD(r)

Single Pore

Figure 10

Drainage in a Capillary Bundle


Suppose we now start with a water wet ( = 0) porous medium - and return to the case
of an ideal parallel bundle of tubes - filled with 100% water (Sw = 1) and then consider
the physics of oil (the non-wetting fluid) displacing this water. For simplicity, we will
take a fully connected capillary bundle of tubes with a uniform PSD and a minimum
radius, Rmin, and a maximum pore size, Rmax (Figure 11). Oil cannot spontaneously
invade water-wet pores and requires an increase in pressure for a displacement to
occur (see earlier discussion).

1
Rmax - Rmin

PSD

f(r)

Rmin

radius, r ->

Rmax
Figure 11

The steps in a Primary Drainage process - and the corresponding drainage capillary
pressures - would be as follows and these are illustrated schematically in Figures
12(c). From the pore occupancies, we calculate the water saturation Sw by summing
the volume of the water-filled pores, divided by the volume of all pores. Similarly, we
may calculate the relative permeabilities (described in a later section).
Step 1: To enter the water filled porous network the oil pressure must be such that Po1
- Pw > Pc,1 = 2/Rmax - this is the minimum entry pressure before the oil can displace
the water from the largest pore where r = Rmax. Figure 12(a). Because Pc,1 is the
smallest of all entry pressures, oil enters the biggest pore first.

14

Petrophysical Input

Step 2: The oil pressure increases such that, Pc1 < Pc2 = 2/r2 where r2 < Rmax. At this
higher capillary pressure, the oil displaces water from all the pores that have Rmax >
r > r2. This leads to a finite oil saturation in the (fully accessible) capillary bundle.
Figure 12(b).
Step 3: The oil pressure increases again such that, Pc1 < Pc2 < Pc3 = 2/r3 where r3 <
r2 < Rmax. At this even higher capillary pressure, the oil displaces more water from all
the pores that have Rmax > r > r3. This leads to a increased oil saturation in the (fully
accessible) capillary bundle. Figure 12(c).
Step 4 (not shown): In principle, for a fully accessible system, we can increase the
capillary pressure to Pcmax = 2/ Rmin and this would displace all the water with 100%
oil (So = 1).
For this Primary Drainage process, the corresponding drainage relative permeabilities
(rel perms) are shown in Figures 13(a) 13(c).
Step 1: At this point, there is only water flow (Sw = 1) and therefore krw = 1 and kro
= 0. Figure 13(a).
Step 2: Now, the water saturation is Sw = Sw2 and krw falls quite rapidly since the water
is now flowing in the smaller pores. Correspondingly, kro rises more rapidly since the
oil is flowing in the larger pores. Figure 13(b).
Step 3: The water saturation is now Sw = Sw3 and krw is relatively low since the water
is now flowing in the smallest pores. Again, kro rises very rapidly since the oil is
flowing in the larger pores as shown in Figure 13(c).
In fact, there are analytical expressions for the relative permeabilities of a capillary
bundle model and these will be introduced later in the course. Also, for a capillary
bundle, the sum of the relative permeabilities turns out to be unity, but this is a result
specific to simple fully accessible models and is not of great importance for real
porous media.

Institute of Petroleum Engineering, Heriot-Watt University

15

(a) Step 1: primary drainage, Pc1 = 2/Rmax

PSD
f(r)

Pc

water

oil

Pc1 = 2
Rmax
Rmin

Rmax

radius, r ->

Sw -->

(b) Step 2: primary drainage, Pc1 < Pc2 = 2/r2 (r2 < Rmax)

PSD
f(r)

Pc

water

oil
r2

Pc2 = 2

r2
Rmin

radius, r ->

Sw -->

Rmax

(c) Step 3: primary drainage, Pc1 < Pc2 < Pc3 = 2/r3 (r3 < r2 < Rmax)

PSD
f(r)

Pc

water

oil

Pc3 = 2

r3

r3

Rmin

radius, r ->

Rmax

Sw -->

Figure 12

16

Petrophysical Input

(a) Step 1: corresponding drainage relative permeabilities

PSD
f(r)

krw

kr

water

kro
Rmin

Rmax

radius, r ->

Sw -->

(b) Step 2: drainage relative permeabilities at Sw = Sw2

krw

PSD
f(r)

water

kr

oil
r2

kro
Rmin

radius, r ->

Sw -->

Rmax

Sw2

(c) Step 3: drainage relative permeabilities at Sw = Sw3

PSD
f(r)

kro

kr

water

oil
krw

r3

Sw -->
Rmin

radius, r ->

Sw3

Rmax

Figure 13

Drainage in connected networks


Drainage physics: We now illustrate what happens when drainage occurs in a
connected network of pores such as that shown in Figure 10(a). Without loss of
generality we can again assume a uniform pore size distribution. The Young-Laplace
equation still applies and the steps shown in Figures 12 and 13 still broadly occur but
there are some important differences from the (fully connected) capillary bundle
model as follows. Although a pore could be occupied by the oil (non-wetting phase)
at a given capillary pressure, Pc1 say where Pc1 =2/r 1, there are two reasons why the
oil may be prevented from invading this target pore:

Institute of Petroleum Engineering, Heriot-Watt University

17

(i) The particular pore may be inaccessible to the invading oil, i.e. the invading oil
phase may not be able to "see" the pore of radius r1, (this would be the case if the
invading oil cluster had not yet reached the target pore). This is the issue of
accessibility and a water filled pore may not be occupied by oil unless the Pc is
above the entry pressure and the oil can access the pore in question;
(ii) The water residing in the target pore could be trapped. Water can be trapped in
a pore when two conditions are satisfied; (a) when. there is no chain of water-filled
pores from the target pore to the outlet of the network (the target pore is then
hydraulically disconnected) and (b) when there are no wetting films coating the
pore surfaces that could allow water to "leak away" from the pore. These films fill
the "corners" of pores as shown for the example of a simple triangular pore in
Figure 14.
water
oil

The drainage process in an interconnected network as described above is illustrated


in Figure 15 and it is known as invasion percolation (with or without trapping). In this
figure, the oil (yellow) is displacing the water (blue) from the left. As noted above,
this is governed by the Young-Laplace equation, but for the oil to displace the water
from a given pore, this pore must also be accessible. There are probably pores in the
large areas of trapped water which are large enough to be occupied by oil but they are
inaccessible.
A sequence of invasion percolation (drainage) calculations in a 2-D network showing
the fluid distributions are shown in Figure16 for co-ordination number, z = 2.667 and
in Figure17 for z = 4 (here, red is oil, blue is water and spaces denote missing pores
i.e. lower co-ordination number). The characteristics of the overall displacement
pattern in these two cases are quite similar but in the lower z case, the initial percolating
capillary finger of non-wetting phase is somewhat more "spindly" or "ramified"
because of lower accessibility. Just at breakthrough, a spanning cluster of non-wetting
phase is formed, i.e. a continuous cluster that goes from the inlet to the outlet. This
spanning cluster has a flowing "backbone" but it also has some dead end branches.
After, breakthrough, the main cluster of non-wetting phase continues to grow and
water is displaced. In the primary drainage simulations shown, water is first displaced
directly - the water is displaced by oil and escapes through a direct route of bulk-filled
pores to the outlet of the network. Later in the flood, at higher oil saturations, the oil
"surrounds" some water filled pores, such that the water appears trapped (see for
example Figures 16 (d) and 17(d)). However, it is evident from the later figures
(Figures 16 (e) and 17(e)) that some of this hydraulically disconnected water has
escaped. This water has escaped through the water films in the corners of the oil-filled
pores (e.g. see the triangular pore in figure above).

18

Figure 14
Invasion percolation where
oil (yellow) displaces the
water

Petrophysical Input

Figure 15
Invasion percolation where
oil (yellow) displaces the
water

Figure 16
Oil/water drainage flood
at various stages in a 2-D
(20 x 20) network for a
water wet system with
coordination number, z =
2.667. This is essentially
invasion percolation with
periodic boundary
conditions (i.e. if the
invading oil leaves the top
of the network, it reenters at the bottom).
Observe that water films
in oil-filled pores are not
visualised

(a) drainage; z = 2.667, So =0.1;

(b) drainage; z = 2.667, So =0.3;

(c) drainage; z = 2.667, So =0.5;

(d) drainage; z = 2.667, So =0.7;

(e) drainage; z = 2.667, So =0.9.

Institute of Petroleum Engineering, Heriot-Watt University

19

(a) drainage; z = 4, So =0.1;

(b) drainage; z = 4, So =0.3;

(c) drainage; z = 4, So =0.5;

(d) drainage; z = 4, So =0.7;

(e) drainage; z = 4, So =0.9.

Accessibility and the accessibility function A(r): The meaning of accessibility and the
definition of the accessibility function, A(r), are shown in Figure 18, together with the
pore filling sequence for a connected network. Compare this with the drainage process
in a fully accessible (capillary bundle) model.

20

Figure 17
Oil/water drainage flood
at various stages in a 2-D
(20 x 20) network for a
water wet system with
coordination number, z =
4. This is essentially
invasion percolation

Petrophysical Input

(a)

p(r)

Accessibility A(r)=A2/(A1+A2)

Pores filled with Hg

M
e
r
c
u
r
y

(b)

10
5

12
2

13
6

7
11

A
i
r

M
e
r
c
u
r
y

10
5

12
2

13
6

7
11

A
i
r

A1
(c)
A2

Figure 18
The idea of accessibility
and the accessibility
function for mercury
(black) invasion into air

Rmin

(d)
Shielded

Rmax

M
e
r
c
u
r
y

10
5

12
2

13

Shielded

7
11

A
i
r

M
e
r
c
u
r
y

10
5

13
6

12
2

7
11

A
i
r

Shielded

The accessibility function is defined as A(r) = A2/(A1+A2) and we can explain this
concept physically as follows..Consider a mercury (Hg) air invasion percolation
process (Figure 18). For the mercury to displace the air from a given pore, this pore
must be accessible. In the sequence of Hg-pore filling in (a) to (d), at certain stages
(e.g. (c)) some pores are shielded, i.e. they are large enough to be invaded but can
not (yet) be seen by the invading mercury. How accessibility is related to the
accessibility function for a drainage displacement is shown in Figure 19(a) 19(c).
where the low occupancy value of the accessibility is seen to be 0 and the high
occupancy is 1 i.e. the phase is at a sufficiently high saturation that it can effectively
see all pores that can be entered at a given (high) capillary pressure.
The underlying theory of the drainage process in an interconnected network is a topic
known as Percolation Theory (See Stauffer and Aharony, Introduction to Percolation
Theory, 2nd Edition, Taylor and Francis, 1992) and we will give a fuller exposition
of this topic later in this chapter.

Institute of Petroleum Engineering, Heriot-Watt University

21

(a) Accessibility in a connected network - above percolation radius, rp


rp

f(r)

Accessibility function

PSD

water

A(r)

A(r) = 0 for
Rmax > r > rp

Rmin

radius, r ->

Rmax

rp

r -->

Rmax

(b) Accessibility in a connected network - just below percolation radius, r < rp


r < rp

PSD
f(r)

water

Not
occupied
= A1

Accessibility function
1
A(r)

A(r) > 0 for


r < rp

Oil occupied
= A2
Rmin

radius, r ->

Rmax

rp

r -->

Rmax

(c) Accessibility in a connected network - well below percolation radius, r << rp


Accessibility function

r << rp

f(r)

PSD

water

oil

Oil occupied
= A2

A(r)

A(r) = 1 for
r << rp

(A1 = 0)
=> A(r) = 1

Rmin

a)

b)

22

radius, r ->

Rmax

r -->

rp

Rmax

Figure 19
The meaning of
accessibility and the
accessibility function,
A(r), in a primary drainage
process

Figure 20
Comparison of a drainage
flood (oil displacing water)
in (a) a 2-D water wet
glass micromodel where
oil is the light fluid and the
etched geometrical pore
pattern can be seen; and
(b) the corresponding 2-D
theoretical network model
calculation of the same
flood where oil is in lighter
blue and the red pores are
water filled (McDougall
and Sorbie, Heriot-Watt
U., unpublished)

Petrophysical Input

A comparison of an experimental drainage flood (oil displacing water) in a glass


micromodel and the corresponding network model calculation is shown in Figure 20.
Figure 20(a) shows the 2D water wet glass micromodel where oil is the light fluid and
the etched geometrical pore pattern can be seen. The corresponding 2D theoretical
network model calculation of the same flood is shown in Figure 20(b) where oil is in
red and the lighter blue pores are water filled. The agreement between these two
figures is sufficiently good to be confident that we have captured the main pore scale
physics of the drainage process in our network model.
3.4 Imbibition at the Pore-Scale

We now consider waterflooding of the same strongly water-wet 2D network models


used in the primary drainage processes described above (i.e. cos = 1 in all pores). A
process where the wetting phase increases such as water injection in this system is
known as imbibition. The pore scale physics of imbibition is not simply the reverse
of the drainage process although there are some features that are common to each. We
noted that in drainage (oil water; or ow), the oil displaced the water by a pistonlike displacement mechanism governed by the Young-Laplace equation. At the pore
level, imbibition - or water displacing oil (wo) in this case - can actually take place
by two distinct mechanisms viz. by piston-like displacement and by snap-off.
Piston-like displacement of oil by water is the reverse of drainage except that it occurs
when one of the phase pressures change such that Po - Pw < Pc = 2/r. The second
mechanism, snap-off, is associated with the flow of wetting phase (water) through
films, which swell around the oil in a pore to form a collar which eventually - at an
appropriate capillary pressure - causes the oil to snap off thus occupying the space with
water. This snap-off process is shown schematically in Figure 21 for a single pore and
in Figure 22 for a 2D interconnected network. A given waterflood will generally
consist of a mixture of the two displacement mechanisms outlined above.
Swelling of wetting phase
to form "collar"

Non-wetting fluid

Figure 21

Wetting fluid

Institute of Petroleum Engineering, Heriot-Watt University

23

(a)

W
a
t
e
r

(b)

10
5

12
2

13
6

O
i
l

11

W
a
t
e
r

10
5

12
2

13
6

7
11

(c)

O
i
l

(d)

Trapped

W
a
t
e
r

10
5

13
6

12
2

7
11

O
i
l

W
a
t
e
r

10
5

13
6

12
2

O
i
l

11

How the snap-off process relates to oil recovery is shown schematically in Figure 23,
which shows a ganglion of oil being snapped-off by water in a water-wet rock. The
oil could only escape through the connected cluster of oil filled pores (since there are
no oil films in a water -wet porous medium). We also note that the isolated blob of
oil left behind in this process in Figure 23 is residual oil since it is trapped and
cannot now move (unless viscous rather than capillary forces are invoked).
Without proof, we note here that the capillary pressure for snap-off is lower than that
for piston-like displacement. Indeed, in a strongly water-wet circular capillary (cos
=1), the snap-off capillary pressure is approximately, Pc = /r i.e. half the value for
piston like displacement. Hence, if a capillary is occupied by oil and the oil pressure
is lowered, the capillary entry pressure for piston-like displacement will be reached
first and - if water is freely available, in adjacent pores for example - then piston-like
displacement will occur first. On the other hand, if the capillary entry pressure drops
below the piston-like entry pressure but no water front is available, then it will not fill
with water. However, if the oil phase pressure, Po, drops sufficiently (or Pw increases
sufficiently) that the snap-off capillary pressure is reached and there are water films
to carry water to that pore, then snap-off will occur. Hence, imbibition is a more
complex process than drainage and the balance between piston-like and snap-off
events that occurs depends on a range of factors such as the range of pore sizes, poregeometry (aspect ratios), the connectivity of the network (z) and the presence/absence
of wetting films (wettability). Clearly, if we have a wide range of pore sizes, then as
we drop the oil phase pressure, Po, the pore that fills next with water is that for which
Po - Pw < Pc; i.e. where the Pc refers to either Pc piston-like in an accessible pore or
snap-off in a smaller but isolated pore which can be supplied wetting phase through
films.

24

Figure 22
Schematic of the snap-off
mechanism in imbibition.
A 3-D view of a pore with
wetting films in the
corners is given in Figure
14

Petrophysical Input

Trapped
Oil

this oil filament is


unstable and "snaps"

continous oil
Flow
oil escapes
through
continous
oil phase

Sand
Grains

Oil Trapping by Filament Snap Off

Trapped
Oil

"snap-off"

continous oil
Flow
oil escapes
through
continous
oil phase

Figure 23
Schematic of snap-off of
an oil ganglion within a
porous medium

Sand
Grains

Oil Trapping by Filament Snap Off

Such a model of imbibition has been implemented in the 2D network of water wet
pores discussed above where we take:
=> Pc for piston-like displacement = 2/r
=> Pc for snap-off events = /r
The phase distribution patterns for imbibition in this network is shown for z = 2.667
in Figures 24(a) - (d) and for z = 4 in Figures 25(a) - (d). Observe that in this case no
direction of flow can be observed, as pores fill simultaneously throughout the
network. In the two cases above, we can determine that the percentages of snap-off
and piston like events in each flood are:
z = 2.667 (Figure 24)
Piston-like = 79 %
z = 4 (Figure 25)
Piston-like = 92.5%

Snap-off = 21 %

Snap-off = 7.5%

Institute of Petroleum Engineering, Heriot-Watt University

25

These results are as we expect since, if we considered a fully accessible capillary


bundle model (as discussed earlier in this section), then only piston like displacements
would occur (for the reasons discussed above). For such a case, the imbibition
capillary pressure for a parallel bundle of tubes would be identical to the drainage
curve i.e. no hysteresis should be observed. For z = 4, the system is moderately
accessible and hence there is only a small fraction of snap-off events (7.5%). For z =
2.667, we find that the fraction of snap-off events grows to 21%. The corollary of this
is that the more snap-off events we observe in the imbibition process, the greater
should be the hysteresis between the primary drainage and the imbibition Pc curves
i.e. the lower the imbibition Pc curve should drop below the primary drainage curve.
This is precisely what is seen in the simulations (Figure 26). Notice that there are more
sources for hysteresis than just snap-off the lower coordination number also leads
to more residual oil trapping and so magnifies the hysteresis effect.
Obviously, we can also derive the relative permeabilities for the drainage and
imbibition processes in the connected networks (see later), which show hysteresis
effects similar to those for the Pc curves of Figure 26.

(a) imbibition z = 2.667, Sw = 0.05;

(c) imbibition z = 2.667, Sw = 0.25;

26

(b) imbibition, z = 2.667, Sw =0.15

(d) imbibition, z = 2.667, Sw =0.281

Figure 24
Oil/water imbuition flood
at various stages in a 2-D
(20 x 20) network for a
water wet system with
coordination number, z =
2.667

Petrophysical Input

Figure 25
Oil/water imbuition flood
at various stages in a 2-D
(20 x 20) network for a
water wet system with
coordination number, z = 4

(a) imbibition z = 4, Sw = 0.05;

(b) imbibition, z = 4, Sw =0.15

(c) imbibition z = 4, Sw = 0.25;

(d) imbibition, z = 4, Sw =0.366

10000

10000
Pc(Pa)

15000

Pc(Pa)

15000

3
2

0000

drainage
0000

1
imbibition
0

0
0

0.25

0.5
Sw

0.75

0.25

0.5
Sw

0.75

Figure 26

In reality, of course, pore geometries in rock samples are far more complicated than
those characterising capillary networks and imbibition is somewhat more complex
(drainage is still relatively straightforward). Micromodel studies by Lenormand and
co-workers (Lenormand and Zarcone, 1984) has uncovered other possible mechanisms,
as illustrated in Figure 27. Their different mechanisms have been termed I1, I2, I3, etc,
where the integer corresponds to the number of pores surrounding a junction that are
filled with nonwetting fluid. We can determine the stability of each meniscus in Figure
27 as the capillary pressure is lowered remember, that the imbibition process is
characterized by increasing meniscus radii as the flood proceeds. Hence, as long as
the next stage of the waterflood results in an increase in meniscus radius, the
displacement remains stable. So, in Figure 27(a) (an I1 mechanism), the displacement
Institute of Petroleum Engineering, Heriot-Watt University

27

begins steadily as the capillary pressure is decreased, with the meniscus moving
gradually from position 1 to position 2 to position 3. However, the next step involves
the meniscus leaving threegrain corners and the radius of curvaturedecreases
this is unstable and the water immediately flows to position 4 and continues to displace
oil from the corresponding pore. Similar reasoning holds for the case shown in
Figure 27(b) (an I2 mechanism) here, however, the meniscus remains stable up to
position 4 and only becomes unstable after grain edge A has been reached and the
meniscus separates. The relative importance of each mechanism during an imbibition
displacement once again depends upon pore geometry, pore size distribution, flowrate
and supply.
x
1

p0

nw

x'

p0
3

01

02

(a)

x' = 2x

x
1
2
5

3
A

1
(c)

5
(b)

(d)

Figure 27

3.5 The Pore Doublet Model

The simplest connected pore system is known as the pore doublet (see Appendix B,
Figure B3). Here, the displacing fluid is introduced to the inlet of the doublet at a
characteristic flowrate q. When the displacing fluid is nonwetting, the widest branch
of the doublet fills first as expected (lowest capillary entry pressure) and wetting phase
is trapped in the thin branch. The situation is far more interesting however when we
consider imbibition in the doublet. A little analysis shows that the frontal velocity in
each branch depends upon the ratio of branch radii (), the supply rate (q), and a
dimensionless number known as the capillary number (Nvc=Lq/R13cos), which
is a ratio of viscous to capillary forces. At low flow rates (small q and small Nvc), the
velocity ratio v2/v1 shows that the thinnest tube fills first and the meniscus in the wider
tube actually retracts. At high rates (large Nvc), we see that v2/v1=R22/R12 and the wider
tube fills first. Fuller treatments of the problem can be found in Moore and Slobod
(1956), Chatzis and Dullien (1983), Laidlaw and Wardlaw (1983), and Sorbie, Wu and
McDougall (1995).
28

Petrophysical Input

3.6 Introduction to Percolation Theory

In the previous two sections, the process of fluid flow in a porous medium has been
considered from the point of view of the fluid. However, it is also possible to consider
the process as being determined by the geometry of the porous medium itself. One
approach, which has since become known as percolation theory, was first used by
Broadbent and Hammersley (1957) to investigate the flow of gas through carbon
granules (for the design of gas masks to be used in coal mines). As well as describing
the flow of fluids and gases through porous media, their theory has subsequently been
used to describe many other diverse processes; such as, the electrical properties of
amorphous semiconductors, the behaviour of crystalline semiconductors containing
impurities, and magnetic phase transitions. Phenomena that are best described by
percolation theory are critical phenomena: characterised by a critical point at which
some property of the system changes abruptly.
In the present context of fluid flow, percolation theory emphasises the topological
aspects of problems, dealing with the connectivity of a very large number of elemental
pores and describing the size and behaviour of connected phase clusters in a welldefined manner.
The primary focus of this section is to discuss how ideas from percolation theory can
be applied more specifically to flow in a porous medium. As a simple, instructive
analogue, consider first a two-dimensional square lattice of capillary elements as
shown in Figure 28(a). The most pertinent concepts from percolation theory will now
be discussed using this geometry. Consider the critical behaviour of the network.
Assume that, initially, all of the tubes are blocked and that they are then opened at
random. For any given geometry there is a unique fraction of tubes that must be open
before flow across the network can commence; this critical fraction is called the
percolation threshold (Pth) and for a simple square lattice has the value Pc=0.5 exactly
(Figure 28(b) shows the distribution of closed and open tubes and Figure28(c) shows
the flowing, spanning cluster of open tubes). One of the most incredible aspects of this
result is that it is independent of the radius distribution; it only depends upon the
topological structure of the network (actually, the co-ordination number (z) and the
Euclidean dimension (d)). In fact, the percolation threshold and system topology are
linked by the equation:

z.Pc =

d
(d 1)

(9)

(see Stauffer and Aharony, 1992). Table B1 in Appendix B shows percolation


thresholds for a variety of two- and three-dimensional geometries.

Institute of Petroleum Engineering, Heriot-Watt University

29

(a)

(b)

(c)

Figure 28

Now, if instead of random opening, the pores are opened systematically beginning
with those of largest radius (top-down filling), it is clear that flow will commence once
a cluster of large open pores spans the system. The radius at which this occurs is known
as the percolation radius, Rp, and is defined implicitly by the equation:

Pth = R Pmax f ( r )dr


R

(10)

where f(r) is the normalised tube radius distribution function and Pc the percolation
threshold. However, the simulation of low-rate drainage processes is carried out using
a top-down invasion percolation model with hydraulic trapping of the wetting phase.
In this case, the injected nonwetting phase first fills the largest pores connected to the
inlet face of the network, and then proceeds along progressively narrower pathways,
sequentially occupying smaller and smaller pores. Although this process appears to
be very different from the pure top-down pore filling, the resulting flowing clusters
are, in fact, identical. Hence, the invasion percolation spanning cluster also appears
at R=Rp.
How does this apply to flow in porous media? Well, many imbibition and drainage
processes exhibit critical behaviour: porous media contain clusters of oil-filled,
water-filled and gas-filled pores and we would only see flow of a particular phase
when the corresponding phase clusters span the system (Figure 29).

30

Petrophysical Input

8 Isolated clusters

3 Isolated clusters +
1 spanning cluster

Figure 29

Moreover, we see no flow below certain critical saturations (Sor, Swi) and these
critical saturations will largely be determined by the connectedness of the porous
medium under investigation (see equation 10 above, where Pth is a function of z).
3.7 A Brief Introduction to Network Modelling of Multiphase Flow

The last section described the concept of percolation theory. The aim of this section
is to broaden the understanding of two-phase flow in porous media; more specifically,
to the simultaneous flow of oil and water through reservoir rock. To this end, concepts
from percolation theory are utilised more fully. The physical porous medium is once
again simulated using a capillary network, but now each element may contain a
different fluid: either water or oil, according to the pore-scale physics of drainage and
imbibition discussed earlier. In the parlance of percolation theory, this type of
analogue is commonly known as a bond percolation model.
The porous medium is initially modelled using a three-dimensional cubic network of
what will be referred to as pore elements; unlike many previous studies, no distinction
is made here between pores and throats. We consider it debatable as to what
constitutes a pore and a throat in a real porous medium, and propose a more abstract
approach based upon effective flow cylinders (the 3Rs approach; McDougall, 1994).
In this formulation, which closely follows the analysis of Heibaet al (1982), theradius
derived from the pore-size distribution is the radius governing the capillary entry
pressure of the pore element and is related to the capillary pressure by:

p c ( r )

1
r

(11)

On the basis of geometrical analysis, Heiba et al postulated an approach for the


estimation of volume and conductance of each pore element in the network. They
reported that the volume, v(r), and conductance, g(r), of each pore element can be
related to the radius governing the pore capillary entry pressure by the following
relationships.

v(r) r

0 3

Institute of Petroleum Engineering, Heriot-Watt University

(12)

31

g(r) r

1 4

(13)

Different combinations of these exponents correspond to different facies; e.g.=3


and=1 would be most appropriate for unconsolidated media, whilst =1 and =4
would apply primarily to consolidated samples. These ideas are central to the 3Rs
approach. One can further dispense with pore/throat arguments by incorporating
effective contact angles into the model formulation (see Dixit et al, 1997). Different
contact angle ranges can be used to mimic the effects of different pore/throat aspect
ratios and the competition between snap-off and pistonlike displacement during
imbibition. In addition, the rich variety of hysteresis phenomena observed during
multiphase flow in consolidated and unconsolidated porous media can be reproduced
and interpreted.
When more than one fluid is flowing through a network, phase occupancy during a
given process (e.g. primary drainage, secondary imbibition etc.) may be characterised
by a set of rules (based upon the physics discussed earlier) which, when combined with
topological considerations (accessibility), give realistic saturation distributions
throughout a displacement. This approach is particularly well-suited to the modelling
of capillary-dominated flow. However, since any increase or decrease in the
saturation of a particular phase depends upon the spatial distribution of that phase, the
computational effort saved in dispensing with unsteady state calculations is more than
offset by the implementation of a clustering algorithm (after Hoshen and Kopelman,
1976). This algorithm is essentially a book keeping exercise which locates and
labels the phase clusters which are distributed throughout the network. These clusters
are continually changing their structure during a displacement, and so the efficacy of
the simulation as a whole is intrinsically linked to the efficacy of the clustering
algorithm itself. A great deal of time and effort has been invested in achieving the
optimum performance of this element of the network simulator. The precise
computational details are dealt with more fully in McDougall (1994).
Although we have rules that tell us how a given displacement proceeds as a function
of capillary pressure, we still need a method of calculating the flow of each phase at
any given stage. In fact, this turns out to be fairly straightforward: we can set a certain
capillary pressure in the model, determine the phase occupancies, effectively freeze
each phase, and use the numerical approach from Section 2.6 to calculate the fractional
flows . This facilitates the calculation of effective permeabilities for the two intertwined
networks (one for each phase) over a range of saturation values. Inherent in this is the
assumption that the flow is stationary, thus steady-state relative permeabilities are
considered here; for work on dynamic relative permeabilities see Blunt and King,
1990. Measured relative permeabilities depend upon the saturation histories, saturation
of the fluids, pore space morphology, wetting characteristics of the fluids, the ratio of
the fluid viscosities and the capillary number. Many of these factors are already
included in the model and will be described later in Section 5.2 when we discuss the
use of network models in calculating capillary pressure curves and relative
permeabilities.
To summarise, network modelling is a powerful tool for increasing our understanding
of multiphase flow in porous media the key element of such an approach lies in
the nterconnected nature of the underlying model. It is particularly well-suited to
32

Petrophysical Input

qualitative sensitivity studies of petrophysical parameters (examining the effects of


connectivity, pore size distribution, dead-end pore space, co-operative filling events,
&c. upon Pc and Krel), although more quantitative studies can also be considered.

f(R)

Imbibition
0.02
0.015
0.01

Pores filled
with wetting
phase

0.005

Pores filled
with nonwetting
phase

20

40

50

R
80

100

80

100

f(R)

Drainage
0.02
0.015
z
y

0.01

Pores filled
with wetting
phase

0.005
Nonwetting
phase

R
20

40

50

Figure 30

4 EXPERIMENTAL DETERMINATION OF PETROPHYSICAL DATA


4.1 Laboratory Measurement of Capillary Pressure

There are a number of ways in which capillary pressure can be measured experimentally
and a few will be briefly discussed here almost all rely upon highly specialised
equipment of varying degrees of sophistication.
Mercury Injection Perhaps the most straightforward approach to capillary pressure
measurement utilises a mercury porosimeter (although manual pumps are also
available, Figure 31(a)(iii). Typically, a 1 diameter cylindrical plug of the porous
sample is inserted into a glass penetrometer (essentially a close-fitting glass
container), which is then put into the machine, evacuated and contacted with mercury
at zero pressure. A table of pressure values ranging from 0-60,000psia is then input
into the porosimeter and the pressure of the mercury in contact with the porous plug
is raised sequentially in a number of discrete steps. The volume of mercury penetrating
the sample is measured at each pressure step (once a given equilibration interval has
elapsed) and the resulting mercury-vacuum capillary pressure curve is output to a file.
Mercury extrusion curves can be similarly measured, although the sub-atmospheric
part of the curve is difficult to obtain without experimental artefact. Note, that the
mercury-vacuum curve cannot be used directly in a reservoir simulator it must be
re-scaled appropriately (the procedure for doing this is described in section 5.1).
Porous Diaphragm The porous diaphragm method is shown in Figure 31(a)(ii). A
water-filled porous plug is placed in an oil-filled (or gas-filled) container, with its
lower face in hydraulic contact with a water-wet glass frit (a glass disc containing
Institute of Petroleum Engineering, Heriot-Watt University

33

micron-scale holes). The pressure in the surrounding nonwetting phase (oil or gas) is
then increased incrementally, and the displaced water leaves the container via the frit.
Saturations are measured at each pressure (once the system has equilibrated) and the
oil-water or gas-water Pc curve can subsequently be determined. Note that the
topology of the invading nonwetting clusters could be different from those obtained
during mercury intrusion, as nonwetting fluid is unable to enter the lower face of the
sample. Note also that a residual wetting phase saturation is also measured using this
method, whereas mercury injection ends with the sample completely filled with
mercury (there is no wetting phase in this case, as the sample is initially evacuated).
Unfortunately, the porous diaphragm method is rather slow (and therefore expensive),
as equilibration times for an oil-water system are generally high. This problem can be
ameliorated somewhat by using water/air systems and rescaling the resulting data.
One final point of note relates to the operating range of the experimental equipment.
The maximum possible driving pressure is determined by the displacement pressure
of the glass frit (diaphragm) upon which the sample is placed. Once the displacement
pressure of the diaphragm has been exceeded, the wetting phase is no longer able to leave
the sample and the experiment ends (Figure 31(b)).
Centrifuge Method The centrifuge method is quick but relies on highly specialist
equipment and analytical treatment of the data. The sample is put into a centrifuge
surrounded by a displacing phase (Figure 31(a)(ii)). The centrifuge is spun at different
speeds and the volume of displaced fluid measured on each occasion. Whilst an
average capillary pressure can be inferred corresponding to each rotational speed, the
local capillary pressure and saturation values vary with distance from the centre of
rotation. Hence, some analysis is required to back-out an appropriate average Pc
curve.
Dynamic Method This method is used in conjunction with steady-state relative
permeability measurements (see below). Here, semi-permeable membranes can be
used to measure wetting-phase and non-wetting phase pressures independently at a
number of different fixed fractional flows. Hence, a direct measure of Pc can be
achieved over a range of saturation values.

34

Petrophysical Input

(i)

(ii)

Nitrogen Pressure

Saran Tube
Crude Oil
Neoprene Stopper

H
Scale of
Squared Paper

P
Nickel-Plated
Spring
Core
Kleenex Paper

Seal of
Red Oil

M
N

Ultra-Fine
Fritted Glass
Disk

Brine

Porous Diaphragm

Centrifuge
0-200 psi Pressure Guage
0-2,000 psi Pressure Guage

(iii)

Regulating Valve
Lucite Window

To
Atmosphere

Cylinder

U-Tube
Manometer
Lucite Window

Figure 31a

Mercury Pump

Institute of Petroleum Engineering, Heriot-Watt University

35

Pc
Pc diaphragm

Pc core

100%
0

Operating
range

Figure 31b

4.2 Laboratory Measurement of Relative Permeability

Relative permeabilities can be measured using either steady-state or unsteady-state


methods: unsteady-state measurements can usually be completed in a much shorter
time, and have consequently become the oil industry standard. The question as to
whether such rapid measurements are representative of conditions in the reservoir,
however, is a matter of some debate.
Steady-State Methods
Steady-state methods are characterised by simultaneous injection of the two phases at
a fixed ratio and known flowrates. Steady-state conditions are assumed to have been
reached once the inlet and outlet fluxes of each phase have equilibrated and/or a
constant pressure drop is seen to exist across the sample. This may take many hours
or even days, depending upon the type of material under investigation and the
measurement technique being used. Once equilibration has been reached, Darcys
law can be used for each phase in turn, resulting in a pair of relative permeability values
valid at that particular saturation. The fluid flux ratio is then changed (whilst keeping
the total flowrate constant), yielding a second set of data once a new steady-state has
been achieved. Similar measurements for a number of different flux ratios can then
be used to give a set of relative permeability curves which span the entire saturation
range. Although the time involved in extracting such data is clearly an important
concern, steady-state measurements performed at low rates should be considered most
indicative of reservoir behaviour. Typical laboratory relative permeability curves are
shown in Figures32 and 33. Some of the more popular experimental techniques will
now be discussed.

36

Petrophysical Input

100
kg ,

k0 , Penn State

kg ,

k0 , single-core dynamic

kg , k0 , dispersed feed

Relative permeability, %

80

kg ,

k0 , Hafford technique

kg ,
kg ,

k0 , gas-drive technique
k0 , Hassler method

60

40

20

Core No. 0-2-A


Berea outcrop
K = 120 md
L = 2.30 cm

20

80
100

40
60
Oil saturation, %

Figure 32

100
kg ,

k0 , Penn State

kg ,

k0 , single-core dynamic

kg , k0 , dispersed feed

Relative permeability, %

80

kg ,

k0 , Hafford technique

kg ,
kg ,

k0 , gas-drive technique
k0 , Hassler method

60

40

Core No. 0-2


Berea outcrop
K = 118 md
L = 7.23 cm

20

20

40
60
Oil saturation, %

80

100

Figure 33

Institute of Petroleum Engineering, Heriot-Watt University

37

The Penn-State Method Originally designed by Morse et al (1947), this technique


has recently become one of the most popular. A typical apparatus is shown in Figure
85a and essentially consists of three similar core plugs mounted in a core holder with
a pair of pressure tappings. Only the central plug is considered for measurement
purposes; one of the two outer units acting as a mixer at the inlet, whilst the other serves
to alleviate capillary end effects at the outlet. The three-plug assembly also facilitates
dismantling of the test section for saturation measurements to be carried out. The
method can be used for both liquid-liquid and liquid-gas measurements and has been
used over a wide range of wettability conditions.
The Hassler Method The laboratory apparatus used for this type of steady-state test
is shown in Figure 35d. Semi-permeable membranes are positioned at both the inlet
and outlet ends of the core sample, which serve to separate the test fluids outwith the
sample whilst permitting two-phase flow to take place within it. The importance of
this is that allows the two pressure drops to be regulated independently, enabling
equilibration of the capillary pressure at both ends of the core. This procedure is
designed to provide a uniform saturation distribution throughout the system, and thus
eliminate any capillary end effects.
The Dispersed-Feed Method
This technique was devised by Richardson et al (1952) in an attempt to introduce
the fluids to the test sample in a more appropriate manner. It utilises an upstream
dispersing medium in order to spread the wetting phase uniformly across the face of
the test sample before entering it. The nonwetting phase is injected into radial grooves
that are machined into the downstream end of the dispersing section, at its boundary
with the core plug. A similar concept lies behind the Hafford apparatus, which injects
nonwetting fluid directly into the sample, whilst the wetting phase must first pass
through a central semi-permeable membrane. Again, the idea is to obtain more
realistic injection behaviour.
Richardson et al (1952) have compared a variety of steady-state methods, and
conclude that all four of the techniques outlined above should give very similar
drainage relative permeability-saturation curves (Figures 32 and 33).
Unsteady-State Methods
Although unsteady-state relative permeability measurements can be made much more
rapidly than those requiring steady-state equilibration, analysis of the resulting data
is more difficult and open to a certain degree of ambiguity. In unsteady-state tests, one
phase is displaced directly from the core sample by the injection of another, at a rate
high enough such that capillary pressure remains negligible. An analysis of the
resulting production data, based upon the Buckley-Leverett frontal advance theory
(Buckley and Leverett, 1942), then permits the determination of a relative permeability
ratio (krw/kro). The relevant theory was given by Welge (1952), who demonstrated that:

fo =

38

1
k
1 + o rw
w k ro

(14)

Petrophysical Input

where fo is the fraction of oil in the outlet stream, and the mi are viscosities. This was
later extended by Johnson et al (1959), who developed a technique for calculating the
individual relative permeabilities from the permeability ratio. Several other methods
also exist (Saraf and McCaffery, 1982; Jones and Roszelle, 1978; inter alia). A sample
calculation is shown in Figure 34 note that we only get relative permeability after
breakthrough of the displacing phase.
1.0
krw
kro

0.8

kri

0.6
0.4
0.2
0.0
0.0

Figure 34

0.2

0.4

0.6

0.8

1.0

Sw

In the past, much work has been carried out to compare these unsteady-state results
with their steady-state equivalents. Unfortunately, although the outcomes appear to
be broadly in agreement, there have been a number of exceptions (e.g. Schneider and
Owens, 1970; Owens et al, 1965; Loomis and Crowell, 1962; Archer and Wong,
1971). The limitations of the unsteady-state approach for determining water-oil
relative permeabilities have been discussed more fully by Craig (1971): the main
concerns relate to the high pressure differentials involved (in excess of 50psi), and the
fact that large viscosity ratios are often used in an attempt to extend the range of twophase flow. The general applicability of the resulting data must consequently be
called into question.
Centrifuge Methods
Centrifuge techniques can provide useful results extremely rapidly, with the added
bonus that they are thought to be free of the viscous fingering problems that
accompany many other unsteady-state methods. They are, however, subject to
problems associated with capillary end effect and cannot be used to quantify the
relative permeability of the displacing phase. Such methods involve monitoring the
production from samples that are initially saturated with one or two phases, with
analytical techniques being used to back-out relative permeability values (see Van
Spronsen, 1982, for a fuller account).

Institute of Petroleum Engineering, Heriot-Watt University

39

Gas

(a)

Thermometer
Packing
nut

Copper
orifice
plate

Electrodes

Inlet

(b)
Gas
pressure
guage

Porous end plate

End
section

Test
section

Mixing
section

Oil
pressure

Differential
pressure taps
Bronze
screen

Oil pressure pad

Highly permeable disk

Outlet

Oil

Gas meter
Intlet

Penn-State

Hafford
Oil burette

(c)

Gas meter

Gas-pressure
guage

(d)

Gas

vacuum

Lucite

Core
material

Core
Lucite-mounted
core

Dispersing
section

Oil

Oil burette

Dispersing
section face

Dispersed Feed

Flowmeter

Hassler

The Effects of Flowrate, Viscosity Ratio, and Interfacial Tension.


At sufficiently low flowrates, microscopic flow behaviour should always be capillarydominated if the system has a strong wetting preference. At such rates, relative
permeabilities in the medium saturation range should consequently be independent of
the viscosity ratio, and this conclusion appears to have been borne out by many
subsequent investigations (Leverett et al, 1939, 1941; Wyckoff and Botset, 1936;
Saraf and Fatt, 1967; Levine, 1954). Near the endpoints, however, and in cases where
the wetting phase is flowing only through thick films, there sometimes exists a strong
hydraulic coupling between the two phases. As a result, the nonwetting phase may
experience hydraulic slip and any analysis using Darcys Law becomes invalid. If
Darcys law is applied regardless, however, nonwetting phase relative permeabilities
greater than unity can often be the result: the experimental results of Odeh (1959)
clearly demonstrate this effect (Figure 36a). If high rates are used in relative
permeability measurements, the subsequent flow behaviour will no longer remain
capillary-dominated, and viscous forces will tend to take over. Moreover, if the
40

Figure 35

Petrophysical Input

associated viscosity ratio is high, then the less viscous invading fluid will begin to
finger through the sample and a condition of uniform saturation will be impossible to
achieve. With this in mind, it is clear that such considerations should not be
overlooked when attempting to interpret a wide variety of unsteady-state coreflood data.
The effect of interfacial tension upon relative permeabilities can also be significant.
In cases where the experimental flowrate is high and the interfacial tension is small,
the capillary forces become less significant and slugs of both fluids may begin to flow
through the same network of pores. In fact, as the interfacial tension approaches zero,
the relative permeability curves actually become straight diagonal lines: i.e. the total
effective mobility of the system remains constant over the entire saturation range.
Experimental results showing this effect are reproduced in Figure 36b.
(a)

(b)

240
1.0

???? m=74.5

200

???? m=82.2

.75

???? m=42.0

????

???? m=5.7

100

.5

????

???? m=5.2

.25

???? m=0.9

????

???? m=0.6

50

100

0
0

.25

Sw

Figure 36

????
.5
Sq

.75

1.0

QUESTIONS WE SHOULD NOW BE ABLE TO ANSWER

What is capillary pressure?


At what scale is it most important?
What things affect meniscus curvature?
How do drainage and imbibition processes differ at the pore scale?
Why can the roughness of pore walls be important?
Is flowrate important in laboratory tests and, if so, why?
What are the advantages of network models over other pore-scale models?

Now, how do we use the knowledge weve obtained so far ?

Institute of Petroleum Engineering, Heriot-Watt University

41

5 EMPIRICAL AND THEORETICAL APPROACHES TO GENERATING


PETROPHYSICAL PROPERTIES FOR RESERVOIR SIMULATION
5.1 Methods for Generating Capillary Pressure Curves and Pore Size Distributions

Background
In this section, we will describe a number of approaches to generate capillary pressure
curves for use in reservoir simulation. We often have to undertake a simulation study
without possessing all of the data we would ideally require perhaps we have only
one experimental capillary pressure curve at our disposal and perhaps it comes from
a part of the reservoir not directly related to our current study. We need some way of
inferring reasonable data from those available to us at the time and, although we may
have to make some gross assumptions, we should be able to invoke our knowledge of
flow at the pore-scale to help us.
Let us begin by reminding ourselves of the drainage case (drainage curves are
frequently used in simulators often erroneously). Remember that, in drainage, we
are increasing the pressure difference between the nonwetting and wetting phases. As
Pc increases, the radius of interface curvature decreases and, using a capillary tube
analogue, the Young-Laplace equation tells us that the nonwetting phase begins to
invade the porous medium when Pc>2cos/Rmax i.e. at the so-called displacement
pressure of the sample. As Pc continues to increase, the nonwetting phase invades
progressively smaller pores. Hence, the nonwetting phase saturation gradually
increases with Pc. The resulting plot of Pc vs Sw is the drainage capillary pressure
curve (Figure 37).

Pc

Water wet
sand

Pc

D.P.
Displacement
pressure

%100
0

Funicular
Pendular

Insular
Funicular

0 0
100% w

Figure 37

42

Petrophysical Input

We can now use this discussion to infer the type of capillary pressure curve that may
result from different samples. For example, how would the pore size distribution
affect the curve? We already know that the displacement pressure is affected by Rmax
the largest pore in the sample but it should also be clear that the range of pore
sizes in a sample can greatly affect the shape of the corresponding Pc curve (Figure
88). Flat plateaux indicate samples that are fairly homogeneous at the pore scale,
whilst steeper curves indicate a large variance in the distribution. Moreover, fine
textured rocks with small cemented grains can be expected to exhibit higher capillary
pressures at a given saturation that coarse-textured media also higher displacement
pressures. So, from our basic understanding of capillary pressure at the microscopic
scale, we have been able to infer a great deal about how we would expect capillary
pressure curves to vary at the continuum (macroscopic) scale.

Pc

Water wet
sand

Irreducible water
saturation

3
2
1

D.P.
All same radii

0
100%

w
0

100%
0

Figure 38

Rescaling Mercury Injection Data


One of the most common ways to derive oil-water or gas-oil capillary pressure curves
is to rescale mercury injection data (which is routinely measured and relatively cheap
to obtain). Consider mercury injection into a porous medium containing a large
exterior pore of radius Rextmax. The capillary pressure required for this pore to be
invaded by mercury is given by the Young-Laplace equation:

(Pce )mercury / air =

2( cos )mercury / air


max
R ext

Institute of Petroleum Engineering, Heriot-Watt University

(15)

43

If we now consider the same medium, filled with water, undergoing oil injection, then
the requisite oil-water capillary pressure is now:

(Pce )oil / water =

2( cos )oil / water


max
R ext

(16)

Now, eliminating Rextmax from equations 15 and 16we arrive at the scaling relationship:

( cos )

(Pce )oil / water = (Pce )mercury / air ( cos ) oil / water

(17)

mercury / air

This clearly holds for any Pc-value, and so a complete oil-water capillary pressure
curve can be obtained from the mercury injection data by applying equation 17at each
saturation.
Experience shows that the ratio of interfacial tensions and cosines on the right hand
side of 17(often difficult to measure accurately) should have a value of approximately
6. However, different ratios have sometimes been needed to reconcile experimental
mercury-air and oil-water data from different rock-types (Figure 39 the ratio is
referred to asFactor in these plots). Such differences may be due to interactions
between contact angle and pore geometry.

290

Mercury injection

40

232

30

174

20

Porosity - 23.0%
Permeability - 3.36 md
Factor - 5.8

10
0

116

Limestone core

20

40

100

80

60

58

60

80

0
100

40

20

225.0
Restored state

25

187.5

Mercury injection

20

150.0
Sandstone core
Porosity - 28.1%
Permeability - 1.43 darcys
Factor - 7.5

15

112.5

10

75.0

37.5

20

40

100

80

60

H20

60

80

0
100

40

20

Mercury capiliary pressure, psi

Restored state

50

30
Water/nitrogen capiliary pressure, psi

348

Mercury capiliary pressure, psi

Water/nitrogen capiliary pressure, psi

60

H20
Hg

Liquid saturation, %

Hg
Liquid saturation, %

Figure 39

44

Petrophysical Input

Derivation of Pore Size Distributions


In addition to providing raw data for the derivation of capillary pressure curves,
mercury injection data can also be used to derive so-called pore-size distributions
(PSDs). The theory for deriving PSDs from intrusion data comes from a paper by
Ritter and Drake (1945), which makes a number of simplifying assumptions: (i) the
pores are circular cylinders in shape, and (ii) all pores of a given radius are accessible
to the mercury (i.e. no accessibility issues). By using the Young-Laplace equation to
convert Pc to pore radius, the PSD (D(r)) can be derived from the capillary pressure
data via the relationship:

D( r ) =

Pc dSmercury

r dPc

(18)

In fact, some algebra shows that this can be rewritten as:

dS
D( r ) = air
dr

(19)

The procedure can be tested using the data shown in Table 2 and you are encouraged
to follow the example.
Unfortunately, the distributions that are produced by this method (and therefore by
commercial porosimeters) are not pore size distributions but pore volume distributions.
The spike (very typical) is caused by accessibility effects prevalent during drainage
processes and the volume associated with some large shielded pores can be
incorrectly assigned to small pores (see Figure 40). To get a better idea of the true PSD
(that is the frequency distribution of pore radii), the Ritter and Drake distribution D(r)
should be divided by r2 for each r-value this generally shows that there are far more
smaller pores than D(r)would suggest. That said, the volume-weighted distributions
provided by commercial porosimeters are still useful lithological fingerprints.

Table 2

(Pc) mercury/air
0
5
6
8
9
11
13
16
18
25
50

sair
0
0.97
0.92
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.15

1. Plot the mercury capilliary pressure curve


2. Derive the oil-water curve (cos=40mN/m
for oil-water and 375mN/m for mercury-air)
3. Use Y-L equation to derive Sw vs r plot
(this is the cumulative intrusion curve)
4. Plot PSD (D(r)) using Ritter and Drake
formula (Hint: use Y-L equation to get a
formula for dSw/dr)

Institute of Petroleum Engineering, Heriot-Watt University

45

De

De

(a)

Figure 40

(b)

The Leverett J-Function


The foregoing discussion has shown that Pc curves are clearly affected by the sample
pore-size distribution, interfacial tension, contact angle, and pore structure. Leverett
(1941) developed a dimensionless group called a J-function to account for
these effects. The J-function has become an important tool for capillary pressure
interpolation and extrapolation and increases the utility of a single capillary pressure
curve. It allows us to adapt a single data set to other areas of a reservoir where data may
be unavailable.
We saw in section 3.4 (Capillary Bundle Models) that a representative pore-radius
can be inferred from sample permeability and porosity via the relation:

~
R

(20)

Hence, a dimensionless group can be formed by defining:


1

Pc k 2
J=

cos

(21)

This can be applied at a number of different saturation values at each saturation,


Pc(Sw) is determined and (21) used to find J(Sw). Often, the contact angle is difficult
to ascertain with any accuracy, and the cos term is often set to unity.
Having found J(Sw), capillary pressure curves corresponding to a range of different
permeability and porosity values (and different fluid combinations via the cos
term) can be determined by simply inverting the J-function, viz:

Pc (Sw ) =

46

J(Sw ). cos
k

1
2

(22)

Petrophysical Input

Example J-functions are shown in Figure41 together with a derived set of Pc-curves.
The procedure can also be used to examine the relationship between permeability
(and/or porosity and/or fluid combinations) and water saturation at different capillary
pressure cut-off values in general, at a given Pc, we find that lower k => higher
Sw (Figure 42).

1.5

1.0

Alundum (consolidated)
Leverett (unconsolidated)

0.371
0.419

(Sw) =

Pc

Hawkins

1.1

0.347
0.151
0.18
0.315
0.116
0.114

()

1.2

Capilliary pressure function, (Sw) =

Kinsella

0.9

0.5

Leverett

20

40
60
Liquid saturation, %
(d)

80

100

20

40
60
Liquid saturation, %
(c)

80

100

()
k

30

40
50
60
70
Water saturation, Sw

80

90

100

(Sw) =

20

Pc
cos

Kinsella Shale

10

10 md

50 md

100 md

200 md

500 md

90

160

72

140

63

120

54

100

45

80

36

60

27

81

()

40

18

20

10

20

30

40

50

60

70

80

90

0
100

Water saturation, %

Figure 41

Institute of Petroleum Engineering, Heriot-Watt University

Pc
cos

Oil water capilliary pressure, psi


(reservoir conditions)

Katie

900 md

180
Height above zero capilliary pressure, ft

27

100

Rangely

0.3

After Amyx Bass and Whiting 1960

80

Leduc
Morano

200

12

40
60
Liquid saturation, %
(b)

0.447

0.4

30

15

20

Theoretical limiting
value for regular
packed spheres

0.6

0
0

18

Alundum

0.1

21

0.7

0.2

24

0.8

(Sw) =

()

1.3

Lim
Sw-1(Sw)

Formation
Woodbine
Weber
Moreno
Viking
Deese
Devonian

Pc
cos

Reservoir
Hawkins
Rangely
El Robie
Kinsella
Katie
Leduc

1.4

47

log k log permeability, md

Capillary pressure = 5 psi

Brine saturation S

20 25 30
Porosity

Transition Zones
We conclude this section on capillary pressure curves with a brief discussion of
transition zones. Transition zones are usually described using the water-wet vertical
bead pack idealization shown in Figure 43. A water-saturated bead pack is allowed
to drain under gravity until capillary equilibrium is reached. At the top of the column,
the water is in the form of discrete rings connected via thin wetting films over grains
called the pendular regime. Moving down the column, the rings become larger and
ultimately coalesce the funicular regime. Near the bottom of the pack, water is
continuous.

48

Figure 42

Petrophysical Input

r
r

(a)

(i)

(ii)

(i)

(ii)

(b)
b
b

Figure 43

(iii)

(iv)

Whilst this idealization gives some limited insight into transition zone behaviour, a
more enlightening picture emerges if we use our knowledge of drainage capillary
pressure at the pore scale this can give us a mental picture of fluid distributions as
we move through a transition zone. Referring to Figure 44a, we see that the
hydrostatic oil and water gradients meet at the OWC (Po=Pw). Above this contact,
Po>Pw and capillary pressure increases as we go further up the reservoir. Figure 44b
shows how the phases would be distributed in the pore network, with oil present in
smaller and smaller pores as we move up through the transition zone.

Institute of Petroleum Engineering, Heriot-Watt University

49

Exploration
Well

Water

Oil

Gas
GOC 5200'
Test Results
at 5250 ft
po = 2402 psia
dpo
= 0.35 psi/ft
dD

Oil
OWC 5500'
Water
Pressure (psia)
2250
5000

2375
2265

Cappilary
Pressure
Increasing

2500

2369

Pcgo

Depth
(feet)

GOC: po = pg = 2385
5250

Pcow
5500

OWC: po = pw = 2490

(a)

Figure 44

(b)

Capillary pressure data can also be used to infer fluid saturations as a function of depth
within a transition zone. Assuming that oil and water pressures remain continuous
throughout the entire height of the zone (doubtful under some circumstances), we can
write equations for each hydraulic gradient (z measures height above OWC and
increases vertically upwards):

dPw
= g w
dz

(23)

dPo
= g o
dz

(24)

Subtracting (23) from (24) leads to:

dPo dPw dPc

=
= g(w o )
dz
dz
dz
which can be integrated to give:

50

(25)

Petrophysical Input

Pc = g(w o )z

(27)

Hence, any capillary pressure curve can be re-plotted to give saturation with depth
information. An example is shown in Figure 45.

830
840
Minimum of 22% connate water
850
860
870
880
Data derrived from
capillary pressure
Data obtained from
seismic logs

Depth, M below sea level

890
900
910
920
930
940
950
960
970
980

Approximate gas-oil contact


990
1000
1010
1020

Figure 45

10

20

30

40

50

60

70

80

90

100

Water saturation scale %

5.2 Methods for Generating Relative Permeabilities

Multiphase flow experiments are costly, difficult and time-consuming to carry out and
it is often infeasible to perform a large number of laboratory sensitivity studies. It
would therefore be highly desirable if a cheap, predictive model for relative permeability
could be developed that only required cheap, easily-accessible data as inputs. In fact,
a number of models are already currently available (although some still lack a degree
of refinement) and we will discuss three such approaches below.
Purcell Model
The Purcell model seeks to utilise cheap capillary pressure data in order to predict
more expensive relative permeabilities. The model revisits the capillary bundle
model and initially uses a little algebra to develop a predictor for absolute permeability.
Institute of Petroleum Engineering, Heriot-Watt University

51

The development is shown in Figure B4 in Appendix B.


Extending the approach to deal with relative permeabilities is relatively straightforward
we simply need to replace the summation in (equation B6 with integrals over water
saturation. The wetting phase relative permeability integral runs from 0 to Swt (the
wetting phase saturation of interest), whilst the nonwetting integral runs from Swt to
1. The equations to be applied at successive values of water saturation are given in
Figure B5.
A graphical recipe for relative permeability prediction via capillary pressure data is
essentially the following:
(i) Obtain capillary pressure data Pc(Sw)
(ii) Plot the curve 1/Pc2 vs Sw (see Figure 46)
(iii)Find the area under the entire curve (=k)
(iv) For any chosen value of water saturation (Swt, say), calculate the area under the
1/Pc2 curve from 0 to Swt. This gives kwt
(v) For the same value of water saturation, calculate the area under the 1/Pc2 curve
from Swt to 1. This gives knwt

14

0.56

12

0.48

Capillary pressure, Pc, atm

10

0.40

Pc

0.32

1
(Pc)2

0.24

0.16

0.08

0
100

1/(capillary pressure)2, 1/Pc2, atm-2

(vi) Use the values obtained from (iii) (v) to determine relative permeabilities

0
80
60
40
20
Percent of total pore space occupied by mercury

Figure 46

There is one clear drawback using this approach, however: the relative permeabilities
add up to unity over the full saturation range and are therefore completely symmetrical.
This limitation led Burdine to add the following improvement.

52

Petrophysical Input

Burdine Model
The Burdine model simply takes the Purcell model and re-scales the endpoints
through the prefactors:

w (Sw ) =

Sw Swi
1 Swi Sor

nw (Sw ) =

1 Sw Sor
1 Swi Sor

(27)

The curves are therefore defined over the wetting-phase saturation range (1-Sor) to (1Swi). The full equations are given in Appendix B (Figure B6). Although a number of
gross assumptions have been made along the way, the Burdine method offers a cheap
alternative to laboratory measurement of relative permeabilities. It is easily coded
into a spreadsheet and is often carried out by practicing reservoir engineers
experience appears to show that the wetting phase prediction is often fairly good,
whilst the nonwetting curve is less well reproduced.
Brooks and Corey Model
Although the models described above yield relative permeability curves that are
qualitatively reasonable, reproduction of experimental data is only achieved via the
introduction of some form of empiricism. Many studies have subsequently relied
entirely upon empirical curve fitting techniques. One of the most popular empirical
correlations is that due to Corey (1954), who proposed the following:
4
k rw Seff

(28)

k rnw (1 Seff )2 (1 S2eff )

(29)

However, it was soon discovered (not surprisingly) that the exponents in Coreys
original equations would have to be varied in order to fit different materials. They
were consequently generalised by Brooks and Corey (1964) to:

k rw = S(eff2 + 3 )

(30)

k rnw = (1 Seff )2 (1 Seff )( 2 + )

(31)

Seff = ( Pcb Pc )

(32)

( Pc Pcb )

with Pcb representing the breakthrough capillary pressure and the pore size
distribution index. Both of these parameters have to be determined experimentally:
Seff vs Pc data is plotted on a log-log scale and a straight line fitted, the slope gives
and the intercept with Seff=1 is assumed to give Pcb.
Network Models
It is quite apparent from this discussion that no satisfactory predictive model currently
exists which can adequately account for the complex jumble of channels that pervade
a porous medium. One of the most promising developments of recent years, however,
has been the possibility of applying interconnected network models to the study of
Institute of Petroleum Engineering, Heriot-Watt University

53

microscopic flow. A detailed discussion now follows to determine the extent to which
the capillary network model is an analogue for two-phase relative permeability
experiments.
McDougall and Sorbie have developed an approach for predicting relative
permeability based on anchoring to capillary pressure data. They have recently
developed a PC-based software package known as MixWet, which allows a wide
range of multiphase simulations to be undertaken at the pore-scale under a variety of
different wettability conditions the interface is shown in Figure 47. Certain pore
scale parameters required as input to MixWet are estimated using the inverse mercury
capillary pressure curve - which we call the R-plot since 1/Pc ~ R, where R is the pore
radius. The full methodology is explained by McDougall et al (SCA, Edinburgh,
2001).

Figure 47

The model calculates the total flow through two intertwined networks (one for each
phase) over a range of saturation values. Inherent in this is the assumption that the flow
is stationary, thus steady-state relative permeabilities are considered here. In all that
follows, the wetting phase is assumed to be water and the non-wetting phase oil. Each
pore in the network is assigned a capillary entry radius from a distribution. At each
stage of the displacement, the total flow is calculated separately for each of the
54

Petrophysical Input

intertwined networks. The results are then converted into normalised permeabilities
as functions of saturation. Additional details can be found in the earlier discussion
presented in section 2.6.
Primary Displacements Consider first the case of a strongly water-wet rock which
is initially 100% saturated with oil. Several possible displacement mechanisms have
been identified in two dimensions (Lenormand and Zarcone, 1984), one of which is
known as snap-off. When brought into contact with the water phase, a strongly
water-wet rock will spontaneously imbibe the wetting fluid via film flow along
irregularities on the pore surfaces. In effect, the water slowly wets all internal grain
surfaces and is essentially present everywhere in the matrix. As this imbibition
process continues, the thin films begin to swell. Eventually, the thinnest pores will
become completely filled with water and the original oil will be displaced if an escape
route exists (see earlier discussion in section 3.4). This continues with the gradual
filling of progressively wider pores until no further displacement is possible since the
oil phase becomes disconnected. The snap-off mechanism is thought to be the most
prevalent imbibition mechanism governing capillary-dominated waterfloods of
consolidated media (low aspect ratio pores). More pistonlike displacements may
occur in media containing high aspect ratio pores (e.g. unconsolidated beadpacks) but
such systems are not considered here (the competition between snap-off and pistonlike
displacements is discussed more fully in Dixit et al, 1997). A typical set of imbibition
relative permeability curves obtained from 3D network modelling are shown in Figure
48 and are seen to correlate well with experimental observations
1.0

0.8

Kri

0.6

0.4

0.2

Figure 48
Imbibition relative
permeability curves from
pore-scale simulation

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Sw

The simulation of low-rate drainage processes is carried out using an invasion


percolation model with hydraulic trapping of the wetting phase. In this case, the
injected non-wetting phase first fills the largest pores connected to the inlet face of the
network, and then proceeds along progressively narrower pathways, occupying
successively smaller pores (see section3.3). The drainage displacement is terminated
at a pre-determined limiting capillary pressure if an unrealistically high capillary
pressure were applied to the network, no irreducible water saturation would remain
at the end of the flood. This drainage model, first proposed by Chandler et al (1982)
and Wilkinson and Willemson (1983), is basically the displacement mechanism
governing mercury porosimetry experiments. A set of simulated drainage relative
Institute of Petroleum Engineering, Heriot-Watt University

55

permeability curves are shown in Figure 49 and again correlate well with experimental
observation.
1.0

0.8

Kri

0.6

0.4

0.2

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Sw

Secondary Displacement Processes Secondary displacement processes are


displacements carried out on a porous medium which is already partially saturated
with the displacing phase. For example, if a primary drainage experiment is followed
by a waterflood, the process is called secondary imbibition. This cycle exhibits an
hysteresis effect between the primary drainage and secondary imbibition non-wetting
relative permeability curves, but very little variation is evident in the wetting phase
curves and the hysteresis effect is usually considered to be negligible. Various theories
relating to pore size distribution and matrix cementation have been put forward in an
attempt to explain this phenomenon, but network modelling of the various secondary
processes provides important clues as to the real causes of hysteresis in porous media.
Both secondary drainage and secondary imbibition have been studied, but for brevity
only the primary drainage-secondary imbibition scenario will be discussed here. The
simulation is shown in Figure 50. The similarity to experimental curves from
consolidated media is striking and the hysteresis effect is clearly duplicated in the case
of the non-wetting phase. The wetting phase curve shows little sign of deviation which
is also in general agreement with experimental findings. In order to explain this
phenomenon, a step by step analysis of the distribution of invaded pores must be
considered. It is found that the hysteresis effect is due to the different physics
governing imbibition and drainage processes; in particular, the associated issue of
accessibility during drainage (McDougall and Sorbie, 1992). Different hysteresis
patterns are exhibited by unconsolidated media and possible causes of this have been
discussed by Jerauld and Salter (1990). More recently, the full range of experimentallyobserved hysteresis phenomena has been modelled and interpreted by Dixit et al
(1997). Both studies show that pore aspect ratio and the competition between snapoff and pistonlike displacement affect the hysteresis trend, although different
methodologies were used to capture the effects.

56

Figure 49
Drainage relative
permeability curves from
pore-scale simulation

Petrophysical Input

1.0
krwint
Primary Drainage

0.8

Secondary Imbibition
krwsec

Kri

0.6

0.4

0.2

Figure 50
Non-wetting phase
hysteresis after a primary
drainage -> secondary
imbibition cycle

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Sw

Comparison with Experiment Although we have presented a number of idealised


relative permeability simulations, the acid test of such a model is to compare relative
permeability prediction with experiment. This is an on-going exercise but some
preliminary examples are shown in Figure 51 together with the matching inverse
capillary pressure data (R-Plots)..

Institute of Petroleum Engineering, Heriot-Watt University

57

Sample F with best-fit


7

0.9

0.8
0.7

Kro sim

0.6

Krg sim

0.5

Kro expt

0.4

Krg expt

5
R, microns

kr

Comparing Sample F with z3.3nu0.9n0lmb4


1

4
3
2

0.3

0.2
0

0.1

0.2

0.4

0.1
0

0.2

0.4

0.6

0.8

Experimental

Sg compensated

Comparing Sample G with z3.3nu1.2n0lmb3

Analytic

40

0.9

35

0.8
0.7

Kro sim

0.6

Krg sim

0.5

Kro expt

0.4

Krg expt

30
R, microns

kr

0.8

Sample F with best-fit

25
20
25
10

0.3
0.2

0.1

0
0

0.1
0

0.2

0.4

0.6

0.2

0.8

Comparing Sample H with z4nu0.6n0lmb3


90
80

0.8
0.7

Kro sim

0.6

Krg sim

0.5

Kro expt

0.4

Krg expt

0.3
0.2
0.1

R, microns

0.9

0.6

Analytic

70
60
50
40
30
20
10
0
0

0.1
0.4

0.8

Sample H with best-fit


100

0.2

0.6

Experimental

0.4
SHg

Sg compensated

kr

0.6
SHg

0.8

0.2

0.4

0.6

0.8

SHg
Experimental

Analytic

Sg compensated

5.3 Hysteresis Phenomena

We have already seen that both capillary pressure and relative permeability curves
exhibit hysteresis that is, they depend upon saturation history (Figure 52 and 53).
There are a number of possible causes for hysteresis but the three main effects are the
following:
(i) Contact angle hysteresis advancing and receding contact angles differ
(reflected in Pc through the Young-Laplace equation)

58

Figure 51
Gas-oil relative
permeabilities (prediction
vs experimental) and
inverse capillary pressure
data for reservoir samples

Petrophysical Input

(ii) Pore structure hysteresis sloping pore walls mean that pores fill and empty at
different capillary pressures
(iii)Topological hysteresis imbibition and drainage processes are different
topologically (film-flow versus fingered invasion)
There is often such a difference between imbibition and drainage curves (both
capillary pressure and relative permeability) that we must make sure that we are using
the correct set of curves in our reservoir simulation studies (for instance, we should
not be using drainage curves as simulation input when modelling a waterflood in a
water-wet reservoir). Built-in numerical models are available in Eclipse and other
commercial reservoir simulators to account for flow reversals at intermediate
saturations.

Pc

a
b

c
Sw

Pc

Swi

1.0

1.0
Sor

kr

kro
Drainage

Secondary
drainage
Intermediate
path

Imbibition

0.0
0.0

0.0
0.0
Swi

Sw

Sw

1.0

1.0

Figure 52

Institute of Petroleum Engineering, Heriot-Watt University

59

48
Water Wet

Capillary Pressure - Cm of Hg

40
Venango Core VL-2
k = 28.2 md
32

24

16
2

20

40

60

80

100

Water Saturation - percent

Figure 53

WETTABILITY CONCEPTS AND APPLICATIONS

6.1 Introductory Concepts

What is Wettability?
The term wettability refers to the wetting preference of a solid substrate in the presence
of different fluid combinations (liquids and/or gases). We have already seen in Section
3,1 and Appendix A6 that the wetting preference of a solid can be characterized by a
contact angle, as shown in Figure 54. The wetting phase is defined as the fluid that
contacts the solid surface at an angle less that 90o.
= 158
= 35
= 30

= 30

Water

Water
Silica Surface

Isooctane

Isooctane
+
5.7% Isoquinoline

Isoquinoline
= 54

= 30

= 106

= 48
Water
Calcite Surface

60

Naphthenic
Acid

Water

Figure 54

Petrophysical Input

This figure clearly demonstrates an important issue, namely, that the wetting preference
of a rock depends not only upon the fluids involved but also upon the mineralogy of
the rock surface. For instance, we see that, in the presence of isoquinoline, water is
nonwetting on a silica substrate but wetting on a calcite substrate. In cases when a solid
has no wetting preference, (i.e. when the angle separating the two fluid interfaces is
close to 90o) the system is said to be of neutral wettability.
The importance of rock wettability cannot be over-emphasised, as it affects almost all
types of core analysis: capillary pressure, relative permeability, waterflood behaviour,
and electrical properties (see Anderson, 1987, for an excellent series of review
articles). The simple reason for this is that wettability affects the location, flow, and
distribution of fluids in a porous medium hence, most measured petrophysical
properties must be affected. Moreover, we shall see later how some core handling
procedures can drastically alter the wettability state of a core: this would invariably
lead to the measurement of SCAL data inappropriate to the reservoir under investigation.
In most of what has been described in this chapter, the assumption has been made that
the system under consideration was water-wet. Indeed, historically, all reservoirs
were believed to be strongly water-wet and almost all clean sedimentary rocks are in
a water-wet condition. An additional argument for the validity of the water-wet
assumption was the following: the majority of reservoirs were deposited in an aqueous
environment, with oil only migrating at a later time. The rock surfaces were
consequently in constant contact with water and no wettability alterations were
possible as connate water would prevent oil contacting the rock surfaces. However,
Nutting (1934) realised that some producing reservoirs were, in fact, oil-wet (the rock
surface was preferentially wetted by oil in the presence of water) and it is now
generally accepted that water-wet reservoirs are the exception rather than the rule
(Table 3 and 4).

Table 3

Table 4

Water-wet
Intermediate wet
Oil-wet
Strongly oil-wet

Water-wet
Intermediate wet
Oil-wet
Total

Contact
Angle
(degrees)
0 to 75
75 to 105
105 to 180

Contact
Angle
(degrees)
0 to 80
80 to 100
100 to 160
160 to 180

Percent of
Reservoirs
8
12
65
15

Silicate
Reservoirs
13
2
15
30

Carbonate
Reservoirs
2
1
22
25

Total
Reservoirs
15
3
37
55

We now know that wettability is a rather complicated issue (still actively being
researched) and that there are a number of different factors affecting reservoir
wettability, including:

Institute of Petroleum Engineering, Heriot-Watt University

61

(i)

Surface-active compounds in the crude oil. These are generally believed to be


polar compounds (polar head and hydrocarbon tail), mostly prevalent in the
heavier crude fractions resins and asphaltenes that form an organic film
or adsorb onto pore walls

(ii)

Brine chemistry

(iii) Brine salinity


(iv) Brine pH
(v)

The presence of multivalent metal cations (Ca2+, Mg2+, Cu2+, Ni2+, Fe3+)

(vi) Pressure and temperature


(vii) Mineralogy (including clays)

In short, surface chemistry determines the wettability state of a reservoir (or, more
correctly, any given region of a reservoir; as wettability can be expected to vary
spatially and possibly temporallythroughout a reservoir).
Having discovered that most reservoirs are not water-wet, we now have to modify
everything that we have learned so far pore-scale physics, capillary pressure
models, relative permeability models, and network models. However, this is not as
difficult as it may first appear; we already understand drainage and imbibition
processes at the pore scale, we know how to define capillary pressure, and we
appreciate the dynamics underlying relative permeability measurements. The following
discussion should help you apply your previous water-wet knowledge to systems that
are not water-wet.
Pore-Scale Effects
The effect of wettability at the pore-scale is shown in Figure 55. If a rock is water-wet,
we have already seen that there is a tendency for water to reside in the tighter pores
and to form a film over the grain surfaces. Oil (the nonwetting phase) resides in the
larger pores. In this case, the term imbibition a process whereby a wetting phase
displaces a nonwetting phase would refer to the displacement of oil by water. The
term drainage would apply to oil displacing water . In an oil-wet system, however,
the situation is reversed oil now forms a thin film over the grain surfaces and water
fills the larger pores. Consequently, in an oil-wet medium,imbibition refers to the
displacement of water by oil, whilst drainage refers to the displacement of oil by
water. These differences clearly have major implications for waterflooding reservoirs
(if a reservoir is oil-wet, a waterflood is a drainage displacement, oil is flushed from
small pores, and oil production via film-flow also becomes important).

62

Petrophysical Input

Oil

(a) Water
Water

Oil

(b) Water
Figure 55

Water

Oil

Oil

Water

Water

Oil

Rock Grains

Oil

Oil

Water

Water

Oil

Rock Grains

The foregoing discussion has clarified some of the pore-scale physics affecting waterwet and oil-wet media. However, given the complex nature of wettability alterations
in reality, a more realistic representation of wettability at the pore scale may be that
shown in Figure 56. Here, a combination of water-wet and oil-wet pathways exists that
allows film-flow access to the displacing phase and film-flow escape for the displaced
phase. Hence, a waterflood would now consist of a combination of imbibition and
drainage events: water initially imbibing along water-wet pathways (displacing oil
from small pores) and then requiring an overpressure to drain oil from the larger oilwet pores. It is clear that the underlying mineralogy and surface chemistry of the
system would determine the connectivity and topology of the wettability pathways
but how can we determine these pathways in short, how can we quantify the
wettability of a porous medium?

Institute of Petroleum Engineering, Heriot-Watt University

63

Water-Wet

Oil-Wet

6.2 Wettability Classification and Measurement

When we come to define the type of wettability associated with a particular porous
sample, we come up against a minefield of inconsistent terminology. It will be
beneficial, therefore, to give some definitions that will appear periodically throughout
this section.
Wettability can be classed as being uniform or non-uniform as follows:
Uniform the wettability of the entire porespace is the same (100% water-wet, 100%
oil-wet, or 100% intermediate-wet) and the contact angle is essentially the same in
every pore;
Non-Uniform this is more characteristic of hydrocarbon reservoirs. The porespace
exhibits heterogeneous wettability, with variations in wetting from pore to pore
(and possibly within a pore) say 70% water-wet pores and 30% oil-wet pores. We
can introduce 2 subdivisions (Figure 57):
Mixed-wet a certain fraction of the
largest pores are oil-wet (there are valid depositional arguments for how this may
come about).
Fractionally-wet no size preference for oil-wetness (there are valid mineralogical
arguments for this).
Given our knowledge of pore-scale displacements, it is clear that each type of
wettability distribution will yield different capillary pressures and relative permeabilities
(and recoveries), and we should be aware of this when we come to interpret SCAL and
wettability test data.

64

Figure 56

Petrophysical Input

F(R)

Mixed-Wet
0.02

0.015

0.01

Water-Wet

0.005

Oil-Wet
R
20

40

60

80

100

80

100

F(R)

Fractionally-Wet
0.02

0.015

Water-Wet

0.01

0.005

Oil-Wet
R
20

40

60

Figure 57

Wettability Measurement
Core wettability can be determined in a number of ways, although three main
quantitative methods are used most frequently:
(i) Contact angle measurement this measures the wettability of a mineral surface
and is mainly used by specialist researchers
(ii) Amott method this measurement gives valuable information regarding the
extent and connectivity of wettability pathways of a core
(iii)U S Bureau of Mines (USBM) method yields an average wettability
measurement, is less informative, but much faster (and therefore cheaper) than
the Amott method
Table 5 gives some expected values for water-wet, neutral-wet, and oil-wet media
from each test (we will explain each of these shortly).

Institute of Petroleum Engineering, Heriot-Watt University

65

Contact angle
Minimum
Maximum
USBM wettability index
Amott wettability index
Displacement-by-water ratio
Displacement-by-oil ratio
Amott-Harvey wettability index

Water-Wet

Neutrally Wet

Oil-Wet

0
60 to 75
W near 1

60 to 75
105 to 120
W near 0

105 to 120
180
W near -1

Positive
Zero
0.3 / 1.0

Zero
Zero
-0.3 < / < 0.3

Zero
Positive
-1.0 / -0.3

Table 5

Other qualitative methods exist, such as rate-of-imbibition tests, NMR methods and
dye adsorption methods, but these are not routinely undertaken. Let us now examine
the three main methodologies in more detail.
Contact angle method
This is shown in Figure 58 and is mainly applicable to pure fluids and isolated mineral
surfaces. A sessile drop (single surface) or modified sessile drop (two surfaces) can
be used. A drop of fluid (oil, say) is generally placed between the mineral surfaces in
a bath of a second fluid (water). The surfaces are then moved in opposite directions
parallel to one another and advancing and receding contact angles can be determined.
These angles usually differ from one another (hysteresis effect) and this is one
component of the hysteresis observed in capillary pressure and relative permeability
curves. Notice also that the early-time behaviour of the measurement is often
misleading you have to wait for the system to equilibrate. Whilst this technique is
often used in wettability research laboratories, it is not carried out routinely in the
industry.

Crystal

Water

Oil

Water
Advancing
Contact
Angle

Oil

Crystal

180
Curve "E" Kareem
Contact Angle (degrees)

150
Curve "D" San Andres
120

90
Curve "B" Deosol

60

Curve "C" Tertiary Kenai

30

Curve "A" Pure Grade C10


0

200

400

600

800

1000

1200

Age of the oil-mineral interface (hours)

66

1400

1600

1800

Figure 58

Petrophysical Input

Amott Test
The Amott test is slow but informative.
It combines both imbibition and forced displacement and is usually carried out via the
following recipe:
(1) Centrifuge core down to Sor
(2) Immerse in oil and measure volume of oil imbibed after 20 hours (Voi)
(3) Centrifuge the core down to Swi and measure the total oil volume (Vot)
(4) Immerse in water and measure volume of water imbibed after 20 hours (Vwi)
(5) Centrifuge the core down to Sor and measure the total water volume (Vwt)
(6) Calculate the displacement-by-oil ratio (Io) and the displacement-by-water ratio
(Iw), i.e.

Io =

Voi
Vot

Iw =

Vwi
Vwt

If Iw ->1 and Io->0, the core is water-wet


If Io ->1 and Iw->0, the core is oil-wet
If Iw=0 and Io=0, the core is said to be neutrally-wet (neither phase has imbibed)
If Iw>0 and Io>0, the core is said to be heterogeneously-wet (mixed- or fractionally-wet)
Often, the indices are combined to give a single index; the Amott-Harvey Index =IwIo, yielding a number between -1 (oil-wet) and 1 (water-wet). If possible, however,
it is better to have access to both Io and Iw, as these individual indices tell far more than
the single Amott-Harvey index. Unfortunately, there are two main drawbacks to the
Amott test: firstly, it takes a long time for fluids to equilibrate, making the test timeconsuming and costly; and secondly, the ratios used in the calculations are very
sensitive to errors in saturation measurement. A far quicker, but less informative,
measure of wettability is given by the USBM test.
USBM Test
The USBM test uses thermodynamic arguments to ascertain the approximate work
done on the system durig centrifuge displacements. This essentially boils down to
calculating areas under Pc curves (= work done during a displacement) as follows
(Figure 59).

Institute of Petroleum Engineering, Heriot-Watt University

67

Capillary Pressure, PSI

10

Water Wet Log A1/A2 = 0.79

II
A1
0
I
A2

a
-10
0

Average Water Saturation, Percent

Capillary Pressure, PSI

10

100

Oil Wet Log A1/A2 = -0.51

II
A1
0
A2
I
b

-10
0

Average Water Saturation, Percent

100

(1) Core driven down to Swi by centrifuge


(2) Core is centrifuged under brine at incremental speeds until Pc=-10psi (NB. water
pressure greater than oil pressure, so Pc=Po-Pw is negative). Measure expelled
oil volume
(3) Core is centrifuged under oil at incremental speeds until Pc=+10psi. Measure
expelled water volume
(4) Plot the two capillary pressure curves
(5) Calculate the areas under the oil-drive (A1) and water-drive(A2) curves
(6) USBM Index (W) = log(A1/A2)

68

Figure 59

Petrophysical Input

Although W theoratically goes from -infinity to +infinity, -1<W<1 is usually reported


(W>0 indicates some degree of water-wetness, W<0 some degree of oil-wetness).
Note, that other methods exist that combine elements of the Amott and USBM tests
(see the literature for details, an example is shown in Figure 60).
6.3 The Impact of Wettability on Petrophysical Properties

Core Handling
The previous discussion has highlighted the importance of wettability at the porescale and its implications for oil recovery (as demonstrated by the Amott and USBM
tests). We should therefore endeavour to preserve the natural state of a core as much
as possible if we wish to derive SCAL data that has relevance to the reservoir under
consideration. There are a number of issues of which we should be aware: (i)
wettability alterations can occur during drilling due to complex mud chemistry, (ii)
pressure and temperature change as a core is brought to surface, (iii) asphaltenes may
subsequently precipitate and light ends may be lost. Pressure coring can help alleviate
some of these problems but this is not always available.
Uniform wetting-contact angle
(Morrow and McCafferty, 1976)

180

133
oil wet

62
intermediate

0
water wet

Amott-IFP Index
(Cuiec, 1991)

-1

-0.3

-0.1

+0.1

+0.3

+1

slightly neutral slightly


oil wet
water wet

Figure 60

oil wet

intermediate

water wet

Three types of core are generally used in core analysis:


Native-state core
ideal if available as it minimises losses, wrap cores in polyethylene film and foil,
then seal in liquid paraffin
Cleaned core should only be used if reservoir is strongly water-wet (rare).
Cleaning procedure should depend upon the crude oil/brine/rock system under
investigation
Restored-state core Only course of action if wettability has been altered. Clean the
core, flow reservoir fluids in correct order, age the core (1000 hours/40 days) and hope
that reservoir conditions have been re-established. Often, of course, this is not the case
and recoveries and relative permeabilities are badly affected (Figure 61).

Institute of Petroleum Engineering, Heriot-Watt University

69

Contaminated

Swi
k0
452, mO 27.6%

ka
25.8% 780, mO

Cleaned

254, mO 31.8%

24.8% 338, mO

Restored-state

202, mO 29.2%

24.8% 338, mO

1
60

NATIVE CORE
CRUDE OIL
CLEANED CORE
CRUDE OIL
CLEANED CORE
REFINED OIL

.8

Relative Permeability

Oil Recovery, Percent PV

50

40

30

20

.4

.2

10

0
0.01

.6

0.1

1.0

Water Injected, Pore Volumes

10

100

.2

.4
.6
.8
Water Saturation, PV

The following discussion, regarding the sensitivity of a number of petrophysical


properties to wettability, will highlight the importance of proper core handling and
cleaning procedures still further.
The Effect of Wettability Upon Electrical Properties
We begin our discussion of petrophysical parameters with a brief look at Archies
equation (Figure 62). This is routinely used by petrophysicists to determine water
saturation in a formation from wireline log data and the exponent (n) has been
determined experimentally from plugs and has a value of about 2 for water-wet/
cleaned cores. Under non-water-wet conditions, however, the fluids are distributed
differently in the pore-space (fewer water films are also present) and we should not
be surprised to learn that Archies equation breaks down (although it is often used
regardless). Consequently, if n=2 is used regardless of the wettability condition
pertaining in the reservoir, then saturation predictions will be wrong (see Pirson and
Fraser, 1960, for an example of how expensive this assumption has been in the past).
In order to demonstrate how inappropriate it would be to assume n=2 for non-waterwet material, we could take reservoir core and actually measure saturations directly.
When taken together with direct resistivity measurements, we could then use Archies
equation to infer what the appropriate exponent should be over a range of saturation
values. The table in Figure 62 shows just such a case: we see that the exponent is not
even constant, implying the breakdown of the assumptions underpinning the model
itself - so use the Archie equation with care

70

Figure 61

Petrophysical Input

Archie Saturation Exponents as a function of


Saturation for a Conducting Nonwetting Phase
Air/NaCi Solution
Brine
Saturation
n
(% PV)
66.2
1.97
65.1
1.98
63.2
1.92
59.3
2.01
51.4
1.93
43.6
1.99
39.5
2.11
33.9
4.06
30.1
7.50
28.4
8.90

_
Sw n

Rt

Ro

Oil/NaCi Solution
Brine
Saturation
n
(% PV)
2.35
64.1
63.1
2.31
60.2
2.46
55.3
2.37
50.7
2.51
44.2
2.46
40.5
2.61
36.8
2.81
34.3
4.00
33.9
7.15
31.0
9

IR

where:

Figure 62

Sw

brine saturation in the porous medium

Rt

Ro

resistivity of the porous medium at


saturation S w, and
resistivity of the 100%
brine-saturated formation

The Effect of Wettability Upon Capillary Pressure


We have already seen an oil/water interface will become curved in order to balance
the pressure jump across it opposing interfacial tension forces. The pressure jump at
which this balance is attained is given by Laplaces equation:

1 1
pc = po pw = +
r1 r2

(33)

where s is the interfacial tension between the two fluids, and r1 and r2 are the two
principal radii of curvature (see section 3.1). The pressure difference Pc is the
capillary pressure and, in oil-water systems, is conventionally taken to be the pressure
in the oil phase minus the pressure in the water phase. For drainage of a water-wet
circular capillary (radius R) and zero contact angle, this relationship becomes:

pc = po pw =

2
R

(34)

Notice, however, that for water to invade an oil-wet capillary, the capillary pressure
must become negative (i.e. pw must become greater than po) The combination of the
words negative and pressure may at first seem confusing, but the term is merely
an artefact of conventional terminology. The concept of negative capillary pressure
is central to displacements in heterogeneously-wet media.

Institute of Petroleum Engineering, Heriot-Watt University

71

When discussing displacements in porous media of heterogeneous wettability, the


terms imbibition and drainage become somewhat confused. For example, waterfloods
in heterogeneously-wet media may be a combination of conventional imbibition and
drainage processes. When the pore network contains a mixture of water-wet and oilwet pores, the waterflood initially proceeds as an imbibition, with water spontaneously
imbibing along water-wet pathways displacing oil from the smallest pores. Eventually,
however, the water has to be forced into oil-wet pores and the displacement becomes
one of drainage.

Primary Drainage

Swi
(Initial Water)

Pc(+ve)

32

A
CURVE
1.DRAINAGE
2.SPONTANEOUS IMBIBITION
3.FORCED IMBIBTION

24

16
Age + Change Wettability

1
8

100

Sw%

Primary Drainage

Swi
(Initial Water)
Pc(+ve)

Forced Oil Drive


(Secondary Drainage)

BEREA CORE
k = 184.3 md

-8

Water Imbibition
(Spontaneous)

Pc(-ve)
Forced Water Drive

Oil
Imbibition
(Spontaneous)

POINT
A. IRREDUCIBLE WETTING SATURATION
B. ZERO - CAPILLARY-PRESSURE
NONWETTING SATURATION
C. IRREDUCIBLE NONWETTING SATURATION

-16

C
-24
0

20

60

40

100

WATER SATURATION, PERCENT P.V.

100

Sw%

48

-48

WATER WET

OIL WET
-40

CAPILLARY PRESSURE, CM HG

40

CAPILLARY PRESSURE - Cm of Hg

80

VENANGO CORE VL - 2
k = 28.2 md

32

24

16
2

-32

-24

-16
1
-8

2
0

20
40
80
60
WATER SATURATION - PERCENT

16

24

12

16

20
40
60
80
WATER SATURATION, PERCENT

100

Drainage capillary pressure characteristics (after Ref.30)

CAPILLARY PRESSURE - Cm of Hg

32

60

40

100

80

Oil-water capillary pressure characterisitics. Ten-sleep


sandstone. oil-wet rock (after Ref.29). Curve 1 - drainage.
Curve 2 - Imbition.

20

20

OIL SATURATION, PERCENT

Capillary pressure characteristics, strongly water-wet rock.


Curve 1 - Drainage. Curve 2 - Imbition.

CAPILLARY PRESSURE, CM.HG

0
0

100

INTERMEDIATE
WET

1
8

BEREA CORE 2-MO16-1


k - 184.3 md

-8

-16

-24

20
40
50
60
WATER SATURATION - PERCENT

100

Oil-water capillary pressure characteristics,


intermediate wettability. Curve 1 - drainage. Curve 2
- spontaneous imbition. Curve 3 - forced imbition.

72

Figure 63
Experimental capillary
pressures in cores for
various processes
( drainage, imbibition) and
wettability conditions
( water-wet, oil-wet,
intermediate-wet )

Petrophysical Input

So, in general, a negative leg in the capillary pressure curve indicates some fraction
of oil-wet porespace. The concept is shown both schematically and for experimental
data in Figure 63.
The Effect of Wettability Upon Relative Permeability
As wettability controls the distribution of phases within the porespace, it is hardly
surprising that wettability has a majorimpact upon relative permeability curves and
subsequent reservoir performance. We can use our pore-scale modeling knowledge
to help explain the differences seen between the slopes of water-wet and oil- wet
curves (Figure 64). In the water-wet case, water imbibes via the smallest pores in the
system, leaving oil resident in large, fast-flowing pores. Consequently, we would
expect the oil relative permeability curve to decrease slowly with increasing water
saturation and the water relative permeability endpoint to remain low this is exactly
what is observed. Conversely, waterflooding an oil-wet medium should lead to a rapid
decrease in oil relative permeability, together with a high water endpoint once
again, this is what is observed.

RELATIVE PERMEABILITY, PERCENT

100

OIL WET
WATER WET

80

CRAIG'S RULES OF THUMBS FOR CETERMINING WETTABILITY

OIL

Water-Wet

Oil-Wet

Interstitial water saturation

Usually greater than


20 to 25% PV.

Generally less than


15% PV.
Frequently less
than 100%.

Saturation at which oil and water relative


permeabilities are equal.

Greater than 50%


water saturation.

Less than 50% water


saturation.

Relative permeability to water at the


maximum water saturation (i.e.,
floodout): based on the effective oil
permeability at reservoir interstitial water
saturation.

Generally less than


30%

Greater than 50%


and approaching
100%.

WATER

60
OIL
40

WATER
20

0
0

20

40

60

80

100

WATER SATURATION, PERCENT P.V.

Relative permeability, % of air permeability

100

100
Water-wet
reservoir

Oil-wet
reservoir

Oil
Oil
60

60

20
S wi
0
0

Water
40
80
Water saturation %

In water-wet system:
S w mostly > 20%
At point A: kro = k rw ;Sw > 50%
k rw at S or / k ro at W wi < 30%

20

S wi
100

0
0

ter

Wa

40
80
Water saturation %

100

In oil-wet system:
S w < 15%
At point A: kro = k rw ;Sw < 50%
k rw at S or / k ro at S wi> 50%

Influence of wettability on relative permeability; after Fertl, OGJ, 22 May 1978

Figure 64

Institute of Petroleum Engineering, Heriot-Watt University

73

In fact, the key features of such curves were presented by Craig (1971) who indicated
the differences between the two in the form of several rules of thumb (see Table in
Figure 64). Many subsequent experimental studies have agreed with these ideas
(Donaldson and Thomas, 1971; Shankar and Dullien, 1981; inter alia) (Figure 65 ).
However, as always, there are a few exceptions that prove the rule(s) pore geometry
and connectivity can change things quite a lot. Nevertheless, the trends are often useful
[Influence of wettability on relative permeability; after Fertl, OGJ, 22 May 1978]

10

10
2

.9
3
4
5

RELATIVE PERMEABILITY TO OIL

.8

1.0

108

UP TO 49 o

USBN
WETTABILITY

0.649

0.02

0.176

.2

-0.222

2.0

-1.250

10.0

-1.333

0.5

.6

.5

5
5

.3

131

.7

.4

DISPLACED
PHASES

RELATIVE PERMEABILITY, FRACTION %

CORE PERCENT
NO.
SILANE

RELATIVE PERMEABILITY TO WATER

138 o
AND GREATER

DISPLACING
PHASES
0.10

0.05

.2

2
DISPLACING
PHASE

.1

1
Nitrogen

0
10

0.01

20

30

40

50

60

70

0
80
0

74

q
up to 49 o
108 o

Nitrogen

Water

Dioclyl Ether

Nitrogen

131o

Nitrogen

138 and
greater

Heptone, Dodecone

WATER SATURATION, PERCENT

DISPLACED
PHASE
Heptone, Dodecone
Dioclyl Ether

0.2
0.4
0.6
0.8
DISPLACED PHASE SATURATION, FRACTION P.V.

1.0

Figure 65

Petrophysical Input

OIL RECOVERY, PERCENT OIL-IN PLACE

80
CONTACT ANGLE
0o
WATER-WET
47 o
90 oo
138
180 o OIL-WET
0 WOR= 25

70

60

50

40

30

20

10

0.2

0.4

0.6

0.8

10

OIL SATURATION. % P.V.

WATER INJECTED, PORE VOLUMES


100
70
50
30
20
10
6
3

10 20
50 100 200 500 1000 2000 5000
PORE VOLUMES OF FLOOD WATER

Figure 66

The Effect of Wettability Upon Waterflood Performance


Clearly, relative permeability affects waterflooding performance through the fractional
flow equation (as does the viscosity ratio) but what type of wettability distribution
leads to the optimum recovery? In the relatively sparse literature on non-uniform
systems there is much disagreement regarding this question (Figures 66 and 67).

Institute of Petroleum Engineering, Heriot-Watt University

75

OIL RECOVERY AT 24 P.V. THROUGHPUT, % ORIGINAL

70

60

50

OHIO SANDSTONE

70

60

50
WATER-WET

OIL-WET

1.0

0.5

0.0

DISPLACEMENT BY WATER RATIO

1.0

0.5
DISPLACEMENT BY OIL RATIO

RESIDUAL WATER SATURATION, PERCENT P.V.

AMOTT WETTABILITY INDEX

40

20

60

80

100

WEIGHT - PERCENT OIL - WET SAND

Figure 67

1.0
krw 1/0
100% WW
75% WW
50% WW
25% WW
0% WW
krw1
kro2/25
krw1/25
krw2
kro2

0.8

Kri

0.6
0.4
0.2
0.0
0.0

76

0.2

0.4

0.6
SW

0.8

1.0

Figure 68
Relative permeability
curves from mixed-wet
systems: a is the oil-wet
pore fraction

Petrophysical Input

Donaldson et al (1969) and Emery et al (1970) performed waterflood experiments


using core plugs which had been aged in crude for varying periods of time. Both
studies showed that the more water-wet the rock, the more efficient the displacement.
Conversely, Kennedy et al (1955), Amott (1959) and Salathiel (1973) have all shown
that the most efficient recovery takes place at close to neutral conditions. The precise
wettability details are unclear from study to study however and so some sort of
modelling approach is appropriate to understand the issue more fully. In the next
section, we therefore go on to examine how network modelling techniques may be
applied in this context.
6.4 Network Modelling of Wettability Effects

Introduction
The wettability characteristics of a porous medium play a major role in a diverse range
of measurements including: capillary pressure data, relative permeability curves,
electrical conductivity, waterflood recovery efficiency and residual oil saturation.
This section describes the development and implementation of a pore-scale simulator
capable of modelling multiphase flow in porous media of nonuniform wettability.
This has been achieved by explicitly incorporating pore wettability effects into the
steady-state models described earlier.
Results are presented which show how (the fraction of pores which are assigned oilwet characteristics) affects resulting relative permeability curves. These have been
used to calculate waterflood displacement efficiencies for a range of wettability
conditions, and recovery is shown to be maximum at close to neutral conditions.
Moreover, simulated capillary pressure data have demonstrated that standard wettability
tests (such as Amott-Harvey and free imbibition) may give spurious results when the
sample is fractionally-wet in nature (McDougall and Sorbie, 1993a).
Here, attention is restricted to mixed-wet systems. The term mixed wettability was
first introduced by Salathiel (1973) to describe systems where the oil-wet pores
correspond to the largest in the sample, the small pores remaining water-wet. Such
situations may arise when oil migrates to water-wet reservoirs and preferentially fills
the larger interstices. The wettability characteristics of these pores may then be altered
by the adsorption of polar compounds and/or the deposition of organic matter from the
original crude, thereby rendering them oil-wet. Fractional wettability, however, is
generally related to the rock matrix itself and is due to the differences in surface
chemistry of the constituent minerals. Because of these variations, crude oil
components may adsorb onto some pore walls whilst ignoring others. This, in effect,
means that fractionally-wet rock contains oil-wet pores of all sizes. Attention here,
however, will be restricted to mixed-wet systems.
Waterflood Simulation Details
One of the advantages of using microscopic network simulators is that physical
properties can easily be ascribed to each pore individually. Here, the wettability of
each pore is controlled so that some are preferentially wetted by water and others by
oil; the fraction of pores wetted by oil is denoted by . The wetting phase contact angle
is taken to be zero in both oil-wet and water-wet pores, i.e. pore walls are very strongly
wetted by the corresponding wetting phase. In all that follows, the term cluster
refers to any group of connected pores containing the same phase, whilst a spanning
Institute of Petroleum Engineering, Heriot-Watt University

77

cluster is a cluster which spans the network connecting the inlet face of the
network to the outlet face.
Each simulation begins with the network 100% saturated with oil, and water is then
introduced at the inlet face. Each waterflood consists of the following two stages:
(1) Water is first allowed to spontaneously imbibe via film flow but only along
continuous water-wet pathways which have access to the inlet. Accessible
water-wet pores are then assumed to become filled via a snap-off mechanism,
whereby the smallest pores are filled first followed by the next smallest and so
on. The defending oil phase can escape from a pore by draining along a pathway
of oil-filled pores which connect it to the outlet.
(2) Once spontaneous imbibition has ceased, the invading water is over-pressured
(i.e. a negative capillary pressure is applied) and now acts as a nonwetting fluid.
The displacement is modelled using an invasion percolation process, and the
water next fills the largest oil-wet pores connected to either the inlet face of the
network or the invading water cluster. If, at any time during the forced
imbibition, water-wet pores are contacted by the invading cluster, then they are
filled spontaneously if the defending oil can escape. Throughout forced
imbibition, oil may escape in two different ways: either
(a) by draining along a pathway of oil-filled pores which connect it to the outlet,
or
(b) by draining via film flow along a pathway of oil-wet pores to the outlet.
Clustering algorithms (following Hoshen and Kopelman, 1976) have been developed
which permit the labelling of both oil clusters and water clusters as well as clusters of
oil-wet and water-wet pores. Note that the fluid clusters are a dynamic phenomenon,
whilst the wettability clusters remain static during a given process.
The relative permeability curves from mixed-wet networks, computed for a variety of
values, are shown in Figure 68. It is apparent that the oil curve loses curvature and
the water curve gains curvature as the oil-wet pore fraction increases. Furthermore, the
crossover point does not steadily move towards lower water saturations (as is often
supposed). For a between 0 and 0.5, it actually shifts to higher saturations; only when
> 0.5 does it begin to moves back towards lower values.
The precise structure of relative permeability curves plays a vital role in determining
reservoir performance and efficiency. The results described above show that
experiments performed on unrepresentative core samples may yield inaccurate curves
and subsequently lead to incorrect field predictions. The precise effect of reservoir
wettability on waterflood performance is now examined in more detail.
Modelling Waterflood Performance
The relative permeability curves described above can now be used in the conventional
fractional flow equations enabling the construction of a family of fractional flow
curves. Buckley-Leverett analyses can then be carried out to uncover how the
microscopic displacement efficiency is expected to be influenced by the wettability
of the system. The results from the pore-scale simulations are shown in Figure 69 and
78

Petrophysical Input

support the conclusion that optimum recovery occurs at some intermediate wettability
state. Indeed, comparisons with a laboratory study on wettability effects (Figure 70)
are extremely encouraging: the simple rule-based simulator implemented here
reproduces the experimental observations very satisfactorily.

Recovery Efficiency

0.8
20PV

0.7
3PV

0.6
BT

0.5
-20

Figure 69
Recovery efficiency vs %
water-wet pores using a
mixed-wet simulator

20

40

60

80

100

% Water-Wet Pores

Recovery Efficiency

0.8

Figure 70
Experimental observations
from Jadhunandan and
Morrow (1991)

0.7
0.6
0.5

20PV

3PV

0.4
0.3
-1.0

BT

-0.6

-0.2

0.2

0.6

1.0

Amott-Harvey Wettability Index

Current Research
Recent development work in the Institute of Petroleum Engineering at Heriot-Watt
University has focussed upon on a new PC-based Visual C++ mixed-wet simulator.
The advantages of the Visual C++ approach are twofold: (i) PC-based material is far
more accessible to the general user than raw Unix-based research code, and (ii) C++
facilitates the creation of dynamic arrays that can be created on the heap and deleted
at the end of function calls this utilises memory far more efficiently, leading to the
possibility of producing far larger networks than before.
The latest version of the mixed-wet simulator (MixWet) is now fully-functional
(Appendix C) and a large number of new features are available. The full cycle of
primary drainage aging water imbibition water drainage oil imbibition
oil drainage are all included. Aging after primary drainage is controlled via the
Institute of Petroleum Engineering, Heriot-Watt University

79

Wettability Parameters group, which can be used to vary the percentage and size
distribution of oil wet pores. Additional checkboxes are available to turn film flow off
and on, calculate relative permeabilities, change boundary conditions from unidirectional flooding to intrusion from all sides, and calculate NMR T2 signals
automatically. Moreover, the random number seed can be set explicitly by the user in
order to examine different realisations of statistically similar networks. A step-by-step
description of the new interface is given in Appendix C.
7 CONCLUDING REMARKS
We conclude with some final thoughts as to the importance of understanding
multiphase flow at the microscopic scale:

The continuum approach fails to explain a great many observations

A lot of the confusion surrounding petrophysics and petrophysical simulator


input can be cleared up by considering the associated small-scale physics

Remember all displacements ultimately occur pore-by-pore

So, by understanding the controlling physics at the pore-scale, we can look at


ways of improving recovery in the future (IFT reduction, depressurisation,
gravity drainage, something more novel?)

The underlying physics can be complicated and a number of controlling


phenomena are intrinsically coupled (wettability, pore structure, capillarity, etc)

BUT, if we dont attempt to understand petrophysics at a fundamental level, then


we run the risk of ever increasing uncertainty in our reservoir predictions

80

Petrophysical Input

REFERENCES
Baker, L. E., 1988, Three-Phase Relative Permeability Correlations, SPE17369,
Proceedings of the Sixth SPE/DOE Symposium on Enhanced Oil Recovery, Tulsa,
OK, April 1988.
Bartell, F.E. and Osterhof, H.J., 1927,Determination of the Wettability of a Solid
by a Liquid, Ind. Eng. Chem., 19 (11), 1277-1280.
Blunt, M.J., 1999,An Empirical Model for Three-phase Relative Permeability,
SPE56474, Proceedings of the SPE Annual Technical Conference and Exhibition,
Houston, TX, October 1999.
Cuiec, L., 1991, Evaluation of Reservoir Wettability and Its Effects on Oil Recovery,
in Interfacial Phenomena in Oil Recovery, N.R. Morrow, ed., Marcel Dekker, Inc.,
New York City, 319.
Dixit, A.B., McDougall, S.R., Sorbie, K.S. and Buckley, J.S., 1999, Pore-Scale
Modelling of Wettability Effects and their Influence on Oil Recovery,SPE Res. Eval.
and Eng., 2(1), 1-12.
Hui, M.-H. and Blunt, M.J., 2000, Pore-Scale Modeling of Three-Phase Flow and the
Effects of Wettability, SPE59309, Proceedings of the SPE/DOE Improved Oil
Recovery Symposium, Tulsa, OK, April 2000.
Johnson, R.E. and Dettre, R.H., 1993, Wetting of Low Energy Surfaces,
inWettability, Surfactant Science Series, Volume 49, J. C. Berg (editor), Marcel
Dekker, New |York.
Kalaydjian, F. J.-M., 1992, Performance and Analysis of Three-Phase Capillary
Pressure Curves for Drainage and Imbibition in Porous Media, SPE24878, Proceedings
of the 67th Annual Technical Conference and Exhibition of the SPE, Washington,
DC, October 1992.
Landau, L.D. and Lifschitz, E.M., 1958, Statistical Physics, Pergamon, London, 471473.
Li, K. and Firoozabadi, A., 2000, Experimental Study of Wettability Alteration to
Preferentially Gas-Wetting in Porous Media and Its Effects, SPE Res. Eval. and
Eng., 3 (2), 139-149.
Morrow, N.R., 1990, Wettability and Its Effects on Oil Recovery, J. Pet. Tech., 42,
1476-1484.
Morrow, N.R. and McCaffery, F., 1978, Displacement Studies in Uniformly Wetted
Porous Media , inWetting, Spreading and Adhesion, G.F. Padday (ed.), Academic
Press, New York City, 289-319.

Institute of Petroleum Engineering, Heriot-Watt University

81

flren, P.E and Pinczewski, W.V., 1995, Fluid Distributions and Pore-Scale
Displacement Mechanisms in Drainage Dominated Three-Phase Flow, Transport in
Porous Media, 20, 105-133.
Rowlinson, J.S. and Widom, B., 1989, Molecular Theory of Capillarity, Clarendon
Press, Oxford.
Stone, H.L., 1970, Probability Model for Estimating Three-Phase Relative
Permeability, J. Pet. Tech., 20, 214-218.
Stone, H.L., 1973, Estimation of Three-Phase Relative Permeability and Residual
Data, J. Can. Pet. Tech., 12,53-61.
van Dijke, M.I.J., Sorbie K.S. and McDougall, S.R., 2000, A Process-Based
Approach for Capillary Pressure and Relative Permeability Relationships in MixedWet and Fractionally-Wet Systems, SPE59310, Proceedings of the SPE/DOE
Improved Oil Recovery Symposium, Tulsa, OK, April 2000
van Dijke, M.I.J. and Sorbie K.S., 2000, A Probabilistic Model for Three-Phase
Relative Permeabilities in Simple Pore Systems of Heterogeneous Wettability,
Proceedings of ECMOR 7, Baveno, Italy, September 2000.
van Dijke, M.I.J., Sorbie K.S. and McDougall, S.R., 2001a, Saturation-Dependencies
of Three-Phase Relative Permeabilities in Mixed-Wet and Fractionally-Wet Systems,
Adv. Water Resour., 24, 365-384.
Dixit, A. B., McDougall, S. R., Sorbie, K.S. and Buckley, J.S.: Pore-Scale Modeling
of Wettability Effects and their Influence on Oil Recovery, SPE Reservoir Eval. and
Eng., 2 (1), pp. 25-36, February 1999.
Fenwick, D.H. and Blunt, M.J.: 1998, Three-Dimensional Modeling of Three Phase
Imbibition and Drainage, Advances in Water Resources, 21, 121-143.
Hui, M.-H. and Blunt, M.J.: 2000, Pore-Scale Modeling of Three-Phase Flow and the
Effects of Wettability, SPE59309, Proceedings of the SPE/DOE Improved Oil
Recovery Symposium, Tulsa, OK, April 2000
Lerdahl, T.R., FLren, P.E. and Bakke, S.: A Predictive Network Model for ThreePhase Flow in Porous Media, SPE59311, Proceedings of the SPE/DOE Conference
on Improved Oil Recovery, Tulsa, OK, April 2000.
Mani, V. and Mohanty, K.K.: 1997, Effect of Spreading Coefficient on Three-Phase
Flow in Porous Media, J. Colloid and Interface Science, 187, 45.
Mani, V. and Mohanty, K.K.: 1998, Pore-Level Network Modeling of Three-Phase
Capillary Pressure and Relative Permeability Curves, SPE Journal, 3, 238-248.
Moulu, J.-C., Vizika, O., Egermann, P. and Kalaydjian, F.: 1999 A New Three-Phase
Relative Permeability Model for Various Wettability Conditions, SPE56477,

82

Petrophysical Input

Proceedings of the SPE Annual Technical Conference and Exhibition, Houston, TX,
October 1999.
McDougall, S.R., Dixit, A.B. and Sorbie, K.S.: 1996,The Use of Capillarity
Surfaces to Predict Phase Distributions in Mixed-Wet Porous Media, Proceedings of
ECMOR V conference, Loeben, Austria, September 1996.
Flren, P.E. and Pinczewski, W.V.: 1995, Fluid Distributions and Pore-Scale
Displacement Mechanisms in Drainage Dominated Three-Phase Flow, Transport in
Porous Media, 20, 105-133.
Pereira, G.G., Pinczewski, W.V., Chan, D.Y.C., Paterson, L. and FLren, P.E.: 1996,
Pore-Scale Network Model for Drainage-Dominated Three-Phase Flow in Porous
Media, Transport in Porous Media, 24, 167-201.
Pereira, G.G.: 2000, Numerical Pore-Scale Modelling of Three-Phase Fluid Flow:
Comparsion between Simulation and Experiment, Phys. Rev. E., 59, 4229-4242.
WAG Report 6: Water Alternating Gas (WAG) Injection Studies Progress Report
No. 6, Heriot-Watt University, Edinburgh, December 2000.
WAGrep5:Water Alternating Gas (WAG) Injection Studies Progress Report No. 5,
Heriot-Watt University, Edinburgh, June 2000.
WAGrep6:Water Alternating Gas (WAG) Injection Studies Progress Report No. 6,
Heriot-Watt University, Edinburgh, December 2000.
Sohrabi, M., Henderson, G.D., Tehrani, D.H. and Danesh, A.: Visualisation of Oil
Recovery by Water Alternating Gas (WAG) Injection using High Pressure Micromodels
- Water-Wet System, SPE63000, Proceedings of the 2000 SPE Annual Technical
Conference, Dallas TX, October 2000.
From AWR paper
Aleman, M.A. and Slattery, J.C.: 1988, Estimation of three-phase relative
permeabilities, Transport in Porous Media, 3, 111-131.
Aziz, K. and Settari, T.: 1979, Petroleum Reservoir Simulation, Applied Science
Publishers, London.
Baker, L. E.: Three-Phase Relative Permeability Correlations, SPE17369,
Proceedings of the 1988 Sixth SPE/DOE Symposium on Enhanced Oil Recovery,
Tulsa, OK, April 1988.
Baker, L. E.: Three-Phase Relative Permeability of Water-Wet, Intermediate-Wet
and Oil-Wet Sandstone, Proceedings of the 7th European IOR -Symposium, Moscow,
October 1993.
Bear, J.: 1972, Dynamics of fluids in porous media, Elsevier, New York, 1972.

Institute of Petroleum Engineering, Heriot-Watt University

83

Bradford, S.A., Abriola, L.M. and Leij, F.J.: 1997 Wettability Effects on Two- and
Three-Fluid Relative Permeabilities, J. Contaminant Hydrology, 28, 171-191.
Burdine, N.T.: 1953 Relative Permeability Calculations from Pore-Size Distribution
Data, Trans AIME, 198, 71-77.
Blunt, M.J.: An Empirical Model for Three-phase Relative Permeability, SPE56474,
Proceedings of the SPE Annual Technical Conference and Exhibition, Houston, TX,
October 1999.
Corey, A.T., Rathjens, C.H., Henderson, J.H. and Wyllie, M.R.J.: 1956 Three-Phase
Relative Permeability, J. Pet. Tech., 8, 3-5; Trans. AIME, 207, 349-351.
Cuiec, L.: 1991, Evaluation of Reservoir Wettability and Its Effects on Oil Recovery,
in Interfacial Phenomena in Oil Recovery, N.R. Morrow, ed., Marcel Dekker, Inc.,
New York City, 319.
Delshad, M. and Pope, G.A.: 1989, Comparison of Three-Phase Oil Relative
Permeability Models, Transport in Porous Media, 4, 59-83.
DiCarlo, D.A., Sahni, A., and Blunt M.J.: Effect of Wettability on Three-Phase
Relative Permeability, SPE40567, Proceedings of the SPE Annual Technical
Conference and Exhibition, New Orleans, September 1998.
Dixit, A.B., Buckley, J.S., McDougall, S.R. and Sorbie, K.S.: 2000,Empirical
Measures of Wettability in Porous Media and the Relationship Between Them,
Transport in Porous Media, in press.
Fayers, F.J. and Mathews, J.D.: 1984, Evaluation of Normalised Stones Methods for
Estimating Three-Phase Relative Permeabilities, SPE Journal, 20, 224-232.
Fayers, F.J.: Extension of Stones Method 1 and Conditions for Real Characteristics
in Three-Phase Flow, SPE Reservoir Engineering, 4, 437-445, November 1989.
Heiba, A.A., Davis, H.T and Scriven, L.E.: Effect of Wettability on Two-Phase
Relative Permeabilities and Capillary Pressures, SPE12172, Proceedings of the SPE
Annual Technical Conference and Exhibition, San Francisco, October 1983.
Heiba, A.A., Davis, H.T and Scriven, L.E.:Statistical Network Theory of ThreePhase Relative Permeabilities, SPE12690, Proceedings of the 4th DOE/SPE
Symposium on Enhanced Oil Recovery, Tulsa, OK, April 1984.
Hustad, O.S. and Hansen, A.-G.: 1996 A Consistent Formulation for Three-Phase
Relative Permeabilities and Phase Pressures Based on Three Sets of Two-Phase
Data, in RUTH: A Norwegian Research Program on Improved Oil Recovery Program Summary, S.M. Skjaeveland, A. Skauge and L. Hinderacker (Eds.), Norwegian
Petroleum Directorate, Stavanger.
Jerauld, G.R. and Rathmell, J.J: Wettability and Relative Permeability of Prudhoe

84

Petrophysical Input

Bay: A Case Study in Mixed-Wet Reservoirs, SPE Reservoir Engineering, 12, 58,
February 1997.
Jerauld, G.R.: General Three-Phase Relative Permeability Model for Prudhoe Bay,
SPE Reservoir Engineering, 12, 255-263, November 1997.
Kalaydjian, F. J.-M.: Performance and Analysis of Three-Phase Capillary Pressure
Curves for Drainage and Imbibition in Porous Media, SPE24878, Proceedings of the
67th Annual Technical Conference and Exhibition of the SPE, Washington, DC,
October 1992.
Kalaydjian, F. J.-M., Moulu, J.-C., Vizika, O., and Munkerud, P.K.: Three-phase
Flow in Water-Wet Porous Media: Determination of Gas/Oil Relative Permeabilities
Under Various Spreading Conditions, SPE26671, Proceedings of the 68th Annual
Technical Conference and Exhibition of the SPE, Houston, TX, October 1993.
Killough, J.E.:Reservoir Simulation with History-Dependent Saturation Function,
SPE Journal, February 1976; Trans. AIME, 261, 37-48.
Land, C.S.: 1968, Calculation of Imbibition Relative Permeability in Two- and
Three-Phase Flow, SPE Journal, June 1968; Trans. AIME, 243, 149-156.
Larsen, J.A. and Skauge, A.: Methodology for Numerical Simulation with CycleDependent Relative Permeabilities, SPE Journal, 3, 163-173, June 1998.
Leverett, M.S. and Lewis, W.B.: 1941, Steady Flow of Gas-Oil-Water Mixtures
Through Unconsolidated Sands, Trans. AIME, 142, 107.
Mani, V. and Mohanty, K.K.: 1997, Effect of Spreading Coefficient on Three-Phase
Flow in Porous Media, J. Colloid and Interface Science, 187, 45.
Morrow, N.R.: 1990, Wettability and Its Effects on Oil Recovery, J. Pet. Tech., 42,
1476-1484.
Moulu, J.-C., Vizika, O., Kalaydjian, F. and Duquerroix, J.-P.: A New Model for
Three-Phase Relative Permeabilities Based on a Fractal Representation of the Porous
Medium, SPE38891, , Proceedings of the SPE Annual Technical Conference and
Exhibition, San Antonio, TX, October 1997.
Moulu, J.-C., Vizika, O., Egermann, P. and Kalaydjian, F.: A New Three-Phase
Relative Permeability Model for Various Wettability Conditions, SPE56477,
Proceeding of the SPE Annual Technical Conference and Exhibition, Houston, TX,
October 1999.
Naar, J. and Wygal, R.J.: 1961, Three-Phase Imbibition Relative Permeability, SPE
Journal, 1, 254-258, 1961; Trans. AIME, 222, 254-258.
Oak, M.J.: Three-Phase Relative Permeability of Intermediate-Wet Berea Sandstone,
SPE22599, Proceedings of the SPE Annual Technical Meeting and Exhibition,
Dallas, TX, October 1991.
Institute of Petroleum Engineering, Heriot-Watt University

85

Flren, P.E. and Pinczewski, W.V.: 1995, Fluid Distributions and Pore-Scale
Displacement Mechanisms in Drainage Dominated Three-Phase Flow, Transport in
Porous Media, 20, 105-133.
Parker, J.C., Lenhard, R.J. and Kuppusamy, T.: 1987, A Parametric Model for
Constitutive Properties Governing Multiphase Flow in Porous Media, Water Resources
Research, 23, 618-624.
Soll, W.E. and Celia, M.A.: 1993 A modified percolation approach to simulating
three-fluid capillary pressure-saturation relationships, Advances in Water Resources,
16, 107-126.
Stone, H.L.: 1970, Probability Model for Estimating Three-Phase Relative
Permeability, J. Pet. Tech., 20, 214-218.
Stone, H.L.: 1973, Estimation of Three-Phase Relative Permeability and Residual
Data, J. Can. Pet. Tech., 12, 53-61.
Temeng, K.O.: Three-Phase Relative Permeability Model for Arbitrary Wettability
Systems, Proceedings of the 6th European IOR-Symposium, Stavanger, May 1991.
Van Dijke, M.I.J., McDougall, S.R. and Sorbie K.S.: 2000a, Three-Phase Capillary
Pressure and Relative Permeability Relationships in Mixed-wet Systems. Transport
in Porous Media, in press.
Van Dijke, M.I.J., Sorbie K.S. and McDougall, S.R.: 2000b, A Process-Based
Approach for Three-Phase Capillary Pressure and Relative Permeability Relationships
in Mixed-Wet Systems, SPE59310, Proceedings of the SPE/DOE Symposium on
Improved Oil Recovery, Tulsa, OK, April 2000.
Vizika, O. and Lombard, J.-M.: 1996, Wettability and Spreading: Two Key Parameters
in Oil Recovery with Three-Phase Gravity Drainage, SPE Reservoir Engineering,
11, 54-60.
Zhou, D. and Blunt, M.: 1997, Effect of spreading coefficient on the distribution of
light non-aqueous phase liquid in the subsurface, J. Contam. Hydrol. 25,1-1

86

Petrophysical Input

8 APPENDIX A: SOME USEFUL DEFINITIONS AND CONCEPTS


A1Sphere Packings
Sphere packings are very useful for qualitative understanding of pore shape and can
also give some insight into porosity measurements. Examples are shown in Figure
A.1 the cubic packing gives the largest porosity (47.6%) whilst the rhombohedral
packing gives the smallest (26%). Let us analyse a simple packing in more detail.

Case 1

Figure A1

Case 4

Case 2

Case 5

Case 3

Case 6

A2Specific Surface
An important measure characterising the grain surface of a porous medium is known
as the specific surface. This plays an important role in the adsorption capacity of a
sample and affects a number of petrophysical measures, including electrical resistivity,
initial water saturation, and absolute permeability. There are two definitions of
specific surface;
(i) The surface area of the pores per unit volume of solids (Ss), and;
(ii) The surface area of the pores per unit bulk volume (Sv). Mathematical relationships
for (i) and (ii) can be derived by noting that the surface area of a sphere is 4R2.
For this idealized system, we find that Ss=3/R and Sv=/2R hence, the specific
surface of a porous medium is inversely proportional to the (mean) grain size.
A3The Pore Size Distribution
Photomicrographs show that pore structure is extremely complex and we may ask
ourselves whether there is any point in trying to impose conformity by attempting
analysis using idealised pore geometries (eg cylinders, etc.). Although it is possible
to partially reconstruct real pore geometries using microtomographical techniques,
these methods are extremely computer-intensive and we really have no option but to
use idealized geometries if we wish to model petrophysical parameters in a costeffective way.

Institute of Petroleum Engineering, Heriot-Watt University

87

In fact, mercury porosimetry analysis still relies upon such models to define so-called
pore-size distributions probability distribution functions defined for the radius
range (Rmin, Rmax) that act as fingerprints for different rock samples. However, we need
to be careful about what we mean by pore-size. Mercury porosimeters assume pores
to be circular cylinders and experimental pore size distribution functions are derived
under these assumptions (in fact, these distributions are actually volume-weighted
throat-size distributions see notes).
A4 The Coordination Number
One measure of the interconnectedness of pore structure is the so-called co-ordination
number (z): it is defined as the average number of branches meeting at a point (Figure
A2). The co-ordination number plays an important role in determining such things as
breakthrough and residual saturations during multiphase displacements, and is
probably one of the most important parameters governing flow processes at the porescale. Co-ordination numbers can vary greatly from system to system: from z=6 for
a simple cubic sphere packing to z 2.8 for Berea sandstone (Doyen, 1988; Dullien,
1979). This wide variation should clearly be taken into account when attempting to
interpret flow behaviour.

x
z
y

Z=6

Z=4

Z=2

A5 Surface and Interfacial Tension


It is well-documented that the molecules of a liquid are closely bound together by
forces of molecular attraction, which serve to keep it as one cohesive assemblage of
particles. Although these forces of cohesion act to cancel one another in the interior
of the liquid, the situation is somewhat different at the surface: at an air/liquid
interface the cohesive forces of the underlying liquid far exceed those of the
competing air molecules, resulting in a net inward pull. The system then behaves as
if the liquid and air were separated by a uniformly stretched membrane, characterised
by a surface tension (). If, instead of a liquid/air system, a liquid/liquid system is
considered, the tensile force is referred to as interfacial tension and is one of the most
important parameters governing multiphase flow in porous media.
A6 Wettability, Contact Angle and Spreading Phenomena
If a solid surface is contacted by a pair of fluids, one of them will tend to have a greater
affinity for that surface than the other. This phase is identified as the wetting phase,
whilst the other is known as the nonwetting phase. The wetting preference of a flat
solid surface can be quantified by inspecting the contact angle and the associated force
balance (Figure A3). This is due to the fact that the magnitude of the contact angle at
equilibrium is intrinsically linked to the free surface energies of the system via
Youngs equation:
88

Figure A2

Petrophysical Input

S1 S2 = 12 cos
where S1 represents the solid/fluid 1 surface free energy, S2 the solid/fluid 2 surface
free energy, and 12 the fluid-fluid interfacial tension. The value of the contact angle
may lie anywhere from 0o to 180o, and is strongly dependent upon the fluid pair and
surface material involved (Figure A4). If =90o, then s1=s2 and neither fluid is
wetting; the system is then described as neutral.
12
s1
Figure A3

s1 - s2= 12 cos

s2

= 158
= 35
= 30

= 30

Water

Water
Silica Surface

Isooctane

Isooctane
+
5.7% Isoquinoline

Isoquinoline
= 54

= 30

Naphthenic
Acid
= 106

= 48
Water

Water

Calcite Surface

Figure A4

A7Spreading Phenomena
Consideration of the trigonometric term in Youngs equation shows that mechanical
equilibrium between two fluids and a solid surface is only possible if:

S1 S2 < 12
i.e. cos
must always be <1. If this condition is violated, however, the system is unstable and
the wetting phase will spontaneously spread on the solid. Although quantification of
this spreadability in fluid/solid systems is not possible at present (there is no current
technique available for measuring either S1 or S2), this is not a problem if the solid
is replaced by a third fluid or a gas.

Institute of Petroleum Engineering, Heriot-Watt University

89

The spreadability of an oil can be quantified with recourse to a spreading co-efficient,


defined by:

So = wa ( ow + oa )
and so initial spreading occurs if this coefficient is either zero or positive. It is evident
from the above discussion that the spreading characteristics of oil/water/gas systems
must have serious implications for a wide range of hydrocarbon recovery processes.
The effect of the spreading coefficient upon three-phase displacements is dealt with
more fully in the notes.

A8Capillarity
Consider the situation where a liquid is in contact with a glass capillary tube. If the
adhesive forces of the liquid to the glass are greater than the cohesive forces in the
liquid, then the interface will curve upwards towards the tube, forming a meniscus
which intersects the tube wall at an angle (Figure A5). The fact that the meniscus
is curved, means that there is now a non-zero vertical component of surface tension,
which acts to pull liquid up the capillary. This continues, until the vertical component
of surface tension is exactly balanced by the weight of fluid below, i.e. when:

R 2 hg = 2 R cos
Total Upward Force = 2RWA cos

WA

2R

where h is the height of the liquid column, R the capillary radius, and r the density of
the fluid. Although the application of this simplistic example to flow in porous media
may not be immediately obvious, it should serve to demonstrate how surface and
interfacial tension forces can play a crucial role in determining fluid distributions at
the pore scale. The full implications of capillary phenomena are apparent in the notes.

90

Figure A5

Petrophysical Input

9 APPENDIX B: MATHEMATICAL BACKGROUND AND


DERIVATIONS
B1Capillary Bundle Permeability
This can be derived as follows:
Consider the system shown in Figure 1a, which consists of n capillary tubes per unit
cross-sectional area. The length of the system is taken to be L. The flow through a
single cylindrical capillary of radius R is given by Poiseuilles law:

q=

R 4 P
8L

where is the fluid viscosity and P the pressure drop across the tubes. Hence, the
total flow (Q) through the porous medium is:

Q = nq =

nR 4 P
8L

Now, the porosity () of the medium is nR2L/(AL)= nR2 (as cross-sectional area
A=1). Hence,

Q=

R 2 P
8L

Setting this equal to Q given by Darcys Law finally leads to:

R2
D2
k=
=
8
32
An interesting aspect of this simple result is that the quantity (k/)1/2 can be thought
of as a sort of average pore diameter.
B2 Carman-Kozeny Equation
The basic premise of this modelling approach is that particle transit times in the actual
porous medium and the equivalent tortuous rough conduit must be the same. Particle
velocity in a rough conduit (vt) is given by the equation:

vt =

R 2 P
2L t

where Lt is the conduit length and particle velocity through the porous medium (vr)
is given by:

Institute of Petroleum Engineering, Heriot-Watt University

91

vr =

kP
L t

For travel times to be the same in both systems, we require Lt/vt=Lr/vr, and this leads
to the relationship:

k=

R 2H
2T 2

where T is called the tortuosity of the sample (Lt/Lr). We now need to determine a
hydraulic radius. The theory utilises established hydraulic practice, in that the
equivalent conduit is assumed to have a radius, RH which takes the form:

RH =

Volume of pores

=
Surface area of pores (1 )Ss

and hence we can write:

k=

3
2 T 2 (1 )2 Ss2

The permeability can be written In terms of an average grain diameter (Dp) by noting
that, for spherical particles, Ss=6/Dp. Hence,
Lr

T=

Lt
Lr

Lt

vt =

R 2 P
2L t

ur =

ur = vr

vr =

kP
Lr
kP
L r

For travel times to be equal

L t Lr
=
vt vr

92

R
2T 2
2

=>

k=

What do we take for R H ?


Volume of pores

RH =
=
Surface area of pores (1 - )Ss
Ss= specific surface / solid volume
=>

k=

2(1- )2 T 2 Ss2

Figure B1
Carman-Kozeny details

Petrophysical Input

D
B

A
l

R2

Figure B2
Generalised Derivation of
Capillary Pressure Across
a Curved Interface

R1

r2

q2

r2

qw
r1

q1

Figure B3
Pore Doublet Details

r1

Currently accepted values of the percolation thresholds of some two-dimensional networks


Network

pcb

Bc = Zpcb

pcs

Honeycomb
Square
Kagom
Triangular

3
4
4
6

1-2 sin(/18) ~ 0.6527*


1/2*
0.522
2 sin(/18) ~ 0.3473*

1.96
2
2.088
2.84

0.6962
0.5927
0.652
1/2*

*Exact result

Currently accepted values of the percolation thresholds of some three-dimensional networks

Table B1
Percolation Thresholds For
a Variety of Network
Geometries

Network

pcb

Bc = Zpcb

pcs

Diamond
Simple cubic
BCC
FCC

4
6
8
12

0.3886
0.2488
0.1795
0.198

1.55
1.49
1.44
1.43

0.4299
0.3116
0.2464
0.119

Institute of Petroleum Engineering, Heriot-Watt University

93

Figure B4
Purcell Method for
Absolute Permeability
Prediction

S = S wt

dS/ (Pc ) 2
kwt

s=0
= krwt = S =1
k
dS/ (P ) 2

s=0

S =1

dS/ (Pc ) 2

knwt
s = s wt
= krnwt = S =1
k
dS/ (P ) 2

s=0

94

Figure B5
Purcell Extension to
Relative Permeability
Prediction

Petrophysical Input

S = Sw

dS
2

2 S = 0 Pc
k rw (Sw ) = ( w (Sw )) S =1
dS
S= 0 Pc2

S =1

dS
2
2 S = Sw Pc
k rnw (Sw ) = ( nw (Sw )) S =1
dS
S= 0 Pc2

where the prefactorsare defined by:


Figure B6
Burdine Extension to
Purcell Model of Relative
Permeabilities

w ( Sw ) =

Sw Swi
1 Swi Sor

nw (Sw ) =

Institute of Petroleum Engineering, Heriot-Watt University

1 Sw Sor
1 Swi Sor

95

10 APPENDIX C: DETAILS OF THE HERIOT-WATT MIXWET


SIMULATOR
Description of Interface Controls
(1)-(3)

The number of nodes (junctions) in the X- Y- and Z-direction

(4)

Coordination Number (Z): the average number of pore elements meeting


at a node

(5)

Distortion Factor used to distort the network, leading to a distribution of


pore lengths (under development). 0.0 gives a regular cubic (3D) or square
(2D) network. 0.5 is the maximum allowed value

(6)

Fraction of oil-wet pores: this refers to the fraction of the total number of
pores in the network that become oil-wet after primary drainage

(6a)

Theta button used to read in water-wet and oil-wet contact angle ranges

(7)-(8)

The mixed-wet and fractionally-wet radio buttons are mutually exclusive.


Mixed-wet assigns oil-wet characteristics to a fraction of the largest pores
containing oil after primary drainage. Fractionally-wet assigns oil-wet
characteristics to a size-independent fraction of the pores containing oil
after primary drainage

(9)

Primary drainage (PD) checkbox: oil displaces water from a 100% waterwet network. This can also be used to examine other generic 2-phase
incompressible drainage displacements (e.g. gas-oil drainage). Successively
higher (positive) capillary pressures are applied to the system and this
drives the displacement

(10)

Water imbibition (WI) checkbox: water imbibes along water-wet pathways


in the system and snaps-off in the smallest oil-filled pores. Aging will
already have taken place before this part of the cycle. The displacement is
controlled by reducing the pressure in the oil phase (i.e. successively lower
(positive) capillary pressures are applied to the system

(11)

Water drainage (WD) checkbox: successively higher (negative) capillary


pressures are applied to the system and water is forced into successively
narrower oil-wet pores

(12)

Oil imbibition (OI) checkbox: oil imbibes along oil-wet pathways, snappingoff the smallest water-filled pores. This is driven by a gradual reduction in
water pressure and this drives the oil imbibition

(13)

Oil drive (OD) checkbox: oil pressure is increased once again and oil is
forced into water-wet pores

(14)

Random number seed changing this value produces a new network


realisation with the same average properties as others that use the same

96

Petrophysical Input

parameter set. In order to get statistically meaningful results, several runs


should be performed using different random number seeds and the results
averaged
(15)

PSD refers to the pore-size distribution exponent (n), where f(r)~rn. n=0
refers to a Uniform distribution, n=3 gives a Cubic distribution, etc. For a
Log-Uniform distribution, this parameter should be set to n=0.999 (not
n=1, as this leads to a singularity). N=10 gives a truncated normal
distribution

(16)

Volume exponent (n) the volume of a pore element is taken to be


V(r)~r n. n =2 gives cylinders

(17)

Conductivity exponent () the conductance of a pore element is taken


to be G(r)~r l . l =4 gives cylinders

(18)

Rmin is the minimum capillary entry radius of the sample

(19)

Rmax is the maximum capillary entry radius of the sample

(20)

Graphics Options: these radio buttons allow the user to view different
aspects of the simulation

(21)

Calculate Kri 2-phase relative permeabilities are calculated when this


checkbox is checked. An SOR algorithm is used to solve the pressure field
and elemental flows can be subsequently calculated

(22)

Mercury? this checkbox is used to allow invasion of mercury from all


sides of the network instead of unidirectional flooding

(23)

Water Films? if this is checked, then water will leave the system through
thin films. This, in effect, leads to no trapping of the water phase during oil
invasion in a 100% water-wet network. This option is checked when
simulating mercury injection, as the intrusion in this case is essentially
mercury-vacuum

(24)

Oil Films? if this is checked, then oil will leave the system through thin
films. This, in effect, leads to no trapping of the oil phase during water
imbibition in a 100% oil-wet network

(25)

NMR Calculations GroupBox partial T2 signals are calculated if this


is checked. This information is then used to calculate the Accessibility
Function (A(R)). Values for Rho andMag. Density should be left
unchanged at present as the NMR section is still an area of active research

(26)

RUN this button is pressed to set the simulation running

Institute of Petroleum Engineering, Heriot-Watt University

97

(27)

Exit used to leave the package. At present, an error box appears upon
termination: this should simply be cancelled and does not affect
performance

The main output file contains information regarding nonwetting phase saturation
(Snw) (oil or gas), wetting phase relative permeability (krw) (water or oil), nonwetting
phase relative permeability (krnw) (oil or gas), current capillary entry radius (R),
number fraction of pores filled with nonwetting phase (p).
Ultimately, it should be possible to simply import an experimental capillary pressure
data set, interpret this automatically and instantiate a network that would serve as
a pre-anchored numerical representation of an experimental sample. This could then
be used for a wide variety of sensitivity studies.

1
2
3
4
5
6
6a

7
8
20
9
10
11
12
13
14

98

26

15
16

27

17
25
18

21

23

19

22

24

Figure 78
Annotated interface (see
text for details of
functions)

Glossary of Terms

CONTENTS
1

GLOSSARY OF COMMON TERMS AND


CONCEPTS IN RESERVOIR SIMULATION
AND FLOW THROUGH POROUS MEDIA
1.1 Some General Definitions
1.2 Reservoir Fluid Properties
1.3 Single Phase Rock Properties
1.4 Multi-Phase Rock/Fluid Properties
1.5 Wettability and Fluid Displacement
Processes
1.6 Oil Recovery Methods, Waterflood Patterns
and Sweep Efficiency
1.7 Terms Used in Numerical Reservoir
Simulation
1.8 Numerical Solution of the Flow Equations
in Reservoir Simulation
1.9 Pseudo-Isation and Upscaling
1.10 Numerical Simulation of Flow in Fractured
Systems
1.11 Miscellaneous - Vertical Equilibrium,
Miscible Displacement and Dispersion

Glossary of Terms

1 GLOSSARY OF COMMON TERMS AND CONCEPTS IN


RESERVOIR SIMULATION AND FLOW THROUGH POROUS
MEDIA
This glossary is intended for use by the reader as a quick reference to terms used
commonly in reservoir engineering in general and in reservoir simulation in particular.
The student is not expected to work through this from begining to end in a systematic
manner. However, the students should make sure that he or she is quite familiar with
all the technical terms that appear in the main text of this unit. It is hoped that this is
of particular use for distance learning students who may have studied the reservoir
engineering distance leasrning unit some time ago but hopefully it will also be of
use to our residential students.

1.1 Some General Definitions

Oilfield Units volumes in oilfield units are barrels (bbl or B);


1 bbl = 5.615 ft3 or 0.159 m3.
A stock tank barrel (STB) is the same volume defined at some surface standard
conditions (in the stock tank) which are usually 60oF and 14.7 psi.
A reservoir barrel (RB) is the same volume defined at reservoir conditions which
can range from ~ 90oF and 1500 psi for shallow reservoirs to > 350oF and 15,000 psi
for very deep (high temperature - high pressure, HTHP) reservoirs. Note that when
1RB of oil is produced it gives a volume generally less than 1B at the surface since
it loses its gas. (See formation volume factor.)
Oil Types: Dry gas; Wet gas; Gas Condensate; Volatile oil; Black oil; Heavy
(viscous) oil; Tar - see Tables 1 and 2 below.

t
o in
eP

2500

C1

l
bb 80%
Bu

De
w

Po
i

B2
D

2000
40%
20%
1500

u
Liq

m
olu
dV

10%
5%

A2

A1

0%

B3

of Pr
oducti
on

B1

1000

Figure 1
Pressure Temperature
Phase Diagram of a
Reservoir Fluid

Critical
Point

Single Phase
Gas Reservoirs
A

Path

nt

Reservoir Pressure, PSIA

3000

Tc

= 127 F
F

3500

Dew Point
or
Retrograde
Gas-Condensate
Reservoirs

Path of Reservoir Fluid


Flui

Bubble Point
or
Dissolved Gas
Reservoirs

Cricondentherm = 250 F

4000

500
0

50

100

150

200

250

300

350

Reservoir Temperature,
T
F

Institute of Petroleum Engineering, Heriot-Watt University

Reservoir
Fluid

Surface
Appearance

GOR
Range

API
Gravity

Typical Composition, Mole %

C1

C2

C3

C4

C5

C6+

Dry gas

Colourless gas

Almost no liquids

100

Wet gas

Colourless gas -

>100 Mscf/bbl
some clear or
straw-coloured
liquid

60o -70o

96

2.7

0.3

0.5

0.1

0.4

Condensate

Colourless gas significant amounts


of light coloured
liquid

3-100
Mscf/bbl
(900-18000 m3/m3)

50o-70o

87

4.4

2.3

1.7

0.8

3.8

"3000
scf/bbl
(500 m3/m3)

40o-50o

64

7.5

4.7

4.1

3.0

16.7

Volatile or
Brown liquid high shrinkage various yellow, red,
oil
or green hues

Black or
low shrinkage
oil

Dark brown
to black viscous
liquid

100- 2500
scf/bbl
(20 - 450 m3/m3)

30o-40o

49

2.8

1.9

1.6

1.2

43.5

Heavy oil

Black viscous liquid

Almost no gas
in solution

10o-25o

20

3.0

2.0

2.0

12.0

71

Tar

Black substance

No gas
viscosity >
10,000cp

< 10

90+

Component

Black Oil

Volatile Oil

Gas-Condensate

Dry Gas

Gas

C1
C2
C3
C4
C5
C6
C7+

48.83
2.75
1.93
1.60
1.15
1.59
42.15

64.36
7.52
4.74
4.12
2.97
1.38
14.91

87.07
4.39
2.29
1.74
0.83
0.60
3.80

95.85
2.67
0.34
0.52
0.08
0.12
0.42

86.67
7.77
2.95
1.73
0.88
....
....

100.00
225
625
34.3
Greenish
Black

100.00
181.00
2000
50.1
Medium
Orange

100.00
112
18,200
60.8
Light
Straw

100.00
157
105,000
54.7
Water
White

100.00
....
Inf.
....

Mol. Wt. C7+


GOR, SCF/bbl
Tank gravity,
0
API Liquid
color

1.2 Reservoir Fluid Properties

Phase: A chemically homogeneous region of fluid which is separated from another


phase by an interface e.g. oleic (oil) phase, aqueous phase (mainly water), gas phase,
solid phase (rock). There is no particular symbol but frequently subscripted o, w, g;
phases are immiscible.

Table 1
Describing various oil types
from dry gas to tar

Table 2
Mole Composition and
Other Properties of Typical
Single-Phase Reservoir
Fluids

Glossary of Terms

Inter Facial Tension (IFT): The IFT between two phases is a measure of energy
required to create a certain area of the interface. Indeed, the IFT is given in dimensions
which are energy per unit area. The symbol for IFT is and units are dyne/cm in
c.g.s. units and N/m (newtons per m) in S.I. units. For example, if both gas and oil are
present in a reservoir then the gas/oil IFT may be in the range, go ~ 0.1-10 mN/m;
likewise. The oil/water value may be in the range, 0w ~ 15 - 40 mN/m. Note that
numerically 1mN/m = 1dyne/cm.
Component: A single chemical species that may be present in a phase; e.g. in the
aqueous phase there are many components - water (H2O), sodium chloride (NaCl),
dissolved oxygen (O2) etc.; in the oil phase there can be hundreds or even thousands
of components - hydrocarbons based on C1, C2, C3, etc. Some of these oil components
are shown in Table 2.
Viscosity: The viscosity of a fluid is a measure of the (frictional) energy dissipated
when it is in motion resisting an applied shearing force; dimensions [force/area.time]
and units are Pa.s (SI) or poise (metric). The most common unit in oilfield applications
is centiPoise (cP or cp). Typical example are:- water viscosity at standard conditions,
w ~ 1 cP; typical light North Sea oils have o ~ 0.3 - 0.6 cP at reservoir conditions
(T ~ 200oF ; P ~ 4000 - 6000 psi); at reservoir conditions, medium viscosity oils
have o ~ 1 - 6 cP; moderately viscous oils have o ~ 6 - 50 cP; very viscous oils
may have o ~ 50 - 1000s cP and tars may have o ~ up to 10000 cP.
Formation Volume Factor: The factor describing the ratio of volume of a phase
(e.g. oil, water) in the formation (i.e. reservoir at high temperature and pressure)
to that at the surface; symbols Bw, Bo etc. For oil, a typical range for Bo is ~1.1 - 1.3
since, at reservoir conditions, it often contains large amounts of dissolved gas which
is released at surface as the pressure drops and the oil shrinks; oilfield units [reservoir
barrels/stock tank barrel (RB/STB)].
API Gravity (API): Definition =
Gas Solubility Factors (or Solution Gas/Oil Ratios): These factors describe the
volume of gas (usually in standard cubic feet, SCF) per volume of oil (usually stock
tank barrel, STB); symbol, Rso and Rsw; units SCF/STB.
Compressibility: The compressibility (c) of a fluid (oil, gas, water) or rock formation
can be defined in terms of the volume (V) change or density () change with pressure
as follows:

c =

1 V 1
=
V P P

Note that this quantity is normally expressed in units of psi-1.


Typical ranges of compressibilities are presented below (from Craft & Hawkins
(Terry revision), 1991):

Institute of Petroleum Engineering, Heriot-Watt University

Compressibilities (units of 10-6 psi-1)


Formation rock
Water
Undersaturated Oil
Gas at 1000psi
Gas at 5000psi

3
2
5
900
50

10
4
100
1300
200

Compressibilities are used in reservoir engineering for Material Balance


Calculations.
Material Balance Equations: Material Balance applied to a reservoir is simply a
volumetric balance. It may be expressed as an equation which relates:
The quantities of oil, gas and water produced.
The reservoir (average) pressure.
The quantity of water influx (e.g. from the aquifer).
The initial oil and gas content of the reservoir.
Essentially the material balance equations described how the energy of expansion
and influx drive production in the reservoir. If there is a sufficiently low (or zero)
fluid influx, the reservoir pressure will decline. One form of the Material Balance
Equation is given below where each term on the left-hand side described a mechanism
of fluid production (from Craft & Hawkins (Terry revision), 1991):

N .( Bt Bti ) +

c .S + c f
N .m. Bti
.( Bg Bgi ) + (1 + m). N.
N . Bti . w wi
Bgi
1 Swi

= N p . Bt + ( Rp Rsoi ). Bg + Bw .Wp
Where the terms have the following meaning:
N = initial reservoir oil, STB;
Np = cumulative produced oil, STB
Boi = initial oil formation volume factor, bbl/STB
Bo = oil formation volume factor, bbl/STB
Bgi = initial gas formation volume factor, bbl/STB
Bg = gas formation volume factor, bbl/STB
Bw = water formation volume factor, bbl/STB
Rsoi = initial solution gas-oil ratio, SCF/STB
Rp = cumulative produced gas-oil ratio, SCF/STB
Rso = solution gas-oil ratio, SCF/STB
We = water influx into the reservoir, bbl
Wp = cumulative produced water, bbl
cw = water isothermal compressibility, psi-1
cf = formation isothermal compressibility, psi p = change in average reservoir pressure, psi
6

.pp + We

Glossary of Terms

Swi = initial water saturation


m = (Initial hydrocarbon vol. of gas cap)/(Initial hydrocarbon vol. of oil)
In practice the material balance equation is often applied in the linear form of
Havlena and Odeh (J. Pet. Tech., pp896-900, Aug. 1963; ibid, pp815-822, July 1964);
see discussion in Craft & Hawkins (Terry revision, 1991).
In the above formulation of the Material Balance Equation, the various terms have
the following interpretation.
Left-Hand Side of the Material Balance Equation

The following terms account for the expansion of any oil and/or gas zones
that may be present in the reservoir:

N .(Bt Bttii ) +

N.m.Bti
.(Bg Bgi )
Bgi

The following term accounts for the change in void space volume which is
the expansion of the formation and the connate water:

c .S + c f
(1 + m).
). N . Bti . w wi
1 Swi

.pp

The next term is the amount of water influx that has occurred into the
reservoir:
We
Right-Hand Side of the Material Balance Equation

The first term of the RHS represents the production of oil and gas:

= N p . Bt + ( Rp Rsoi ). Bg

The second term of the RHS represents the production of water:


Bw.Wp

1.3 Single Phase Rock Properties

Porosity: the fraction of a rock that is pore space; common symbol, Porosity varies
from 0.25 for a fairly permeable rock down to 0.1 for a very low permeability
rock; there may be an approximate correlation between k and .
Pores & pore throats: The tiny connected passages that exist in permeable rocks;
typically of size 1m to 200 m; they are easily visible in s.e.m. (scanning electron
microscopy). Pores may be lined by diagenetic minerals e.g. clays. The narrower
constrictions between pore bodies are referred to as Pore Throats. See Figure 2:
Institute of Petroleum Engineering, Heriot-Watt University

illite

illite
quartz
quartz

10mm
~1mm

Permeability: The fluid (or gas) conducting capacity of a rock is known as the
permeability; symbol k ; units Darcy (D) or milliDarcy (mD); dimensions -> [L]2.
Permeability is found experimentally using Darcy's Law (see below). Permeability
can be anisotropic and show tensor properties (see below) - denoted k . Probably
the most important quantity from the point of view of the reservoir engineer since
its distribution dictates connectivity and fluid flow in a reservoir. Timmerman (p. 83,
Vol. 1, Practical Reservoir Engineering, 1982) presents the rule:
Classification
poor to fair
moderate
good
very good
excellent

Permeability Range (mD)


1
- 15
15 - 50
50 - 250
250 - 1000
> 1000

k/ Correlations:
It has been found in many systems that there is a relationship
between permeability, k, and porosity, . This is not always the case and much scatter
can be seen in a k/ crossplot. Broadly, higher permeability rocks have a higher
porosity and some of the relationships reported in the literature are shown below.
Some examples of k/ correlations which have appeared in the literature are shown
in Figure 3:

Figure 2

Glossary of Terms

Core
Core
Permeabilty
Permeabilty
(md)
(md)

100.0
100.0
50.0
50.0

10.0
10.0
5.0
5.0

1.0
1.0
0.5
0.5

0.1
0.1
0.05
0.05

Figure 3
Permeability/Porosity
Correlation for Cores from
the Bradford Sandstone

0.01
0.01

6
6

8
8

10
10

12
14
16
12
14
16
Core Porosity (%)
Core Porosity (%)

18
18

20
20

22
22

10,000
10,000

Core
Core
Permeabilty
Permeabilty
(md)
(md)

1,000
1,000

100
100

10
10

Figure 4
Permeability/Porosity
Correlation for Cores from
the Brent Field

1
10
0

10
10

20
20
Core Porosity (%)
Core Porosity (%)

30
30

Darcys Law: Originally a law for single phase flow that relates the total volumetric
flow rate (Q) of a fluid through a porous medium to the pressure gradient (P/x) and
the properties of the fluid ( = viscosity) and the porous medium (k = permeability;
A = cross-sectional area): Note that Darcy's law can be used to define permeability
using the quantities defined in Figure 5 as follows:

Institute of Petroleum Engineering, Heriot-Watt University

k.A P
Q =

x
Note that the in the equation in Figure 5 is a factor for units conversion (see Chapter
2).
Darcy Velocity: This is the velocity, u, calculated as, u = Q/A; this may be expressed
as,

u=

k P
Q
=
A
x

Pore Velocity: This is the fluid velocity, v, given by,

v=

Q
u
=
A.

P
Q

Q
L

Q = .

k.A P

.
L

Permeability Anisotropy: Since permeability can be directional, then it is possible


for kx ky kz in a given system. This is often seen in practice when comparing the
horizontal permeability, (kh), with the vertical permeability, (kv) - usually as the ratio,
(kv/kh). It is often found (kv/kh) < 1, i.e. there is more resistance to vertical flow than
horizontal flow. The value of (kv/kh) must be evaluated with respect to the scale (i.e.
the size) of the sample or system. The value of this quantity will be different in a
core plug or in a large simulator grid block in which the core plug was a small part.
The origin of the anisotropy may be quite different in each case.
At the small (core plug) scale, anisotropy may come from fabric anisotropy at the
grain level or from lamination at the slightly larger scale (laminaset scale). At the
larger scale (grid block), the anisotropy may arise from larger scale heterogeneity,
even though locally the component rock facies are completely isotropic. This is
illustrated in Figure 6.

10

Figure 5
Schematic of the Single
Phase Darcy Law

Glossary of Terms

Core Plug

Small Scale

Fabric Anistropy

kh
kv
Hi k lamina

Rock Grains

Lo k lamina

Lamination

Grid Block Scale (100s m)


kv

kh

Figure 6
Permeability anistropy at
different scales

For Whole System

Heterogeneity
Anistropy
Low Perm Lenses
or Shales
(kv/kh) = 1
Low Perm Sand
(kv/kh) = 1

1.4 Multi-Phase Rock/Fluid Properties

Saturation: The saturation of a phase (oil, water, gas) is the fraction of the pore
space that it occupies (not of the total rock + pore space volume); symbols Sw, So
and Sg ; saturation is a fraction, where Sw + So + Sg = 1. Multi-phase flow functions
such as relative permeability and capillary pressure (see below) depend strongly on
the local fluid saturations.
Residual Saturation: The residual saturation of a phase is the amount of that phase
(fraction pore space) that is trapped or is irreducible; e.g. after many pore volumes of
water displace oil from a rock, we reach residual oil saturation, Sor; the corresponding
connate (irreducible) water level is Swc (or Swi); the related trapped gas saturation is
Srg; at the residual or trapped phase saturation the corresponding relative permeability
(see below) of that phase is zero. Strictly, we should refer to the phases in terms of
wetting and non-wetting phases - the residual phase of non-wetting phase is trapped
in the pores by capillary forces. Typically, in a moderately water wet sandstone, Sor
~ 0.2 - 0.35. The amount of trapped or residual phase depends on the permeability
and wettability of the rock and a large amount of industry data is available on this
quantity: For example, the relationship for k vs. Swc (or Swi) is shown for a range of
reservoir formations, Figure 7.

Institute of Petroleum Engineering, Heriot-Watt University

11

10,000
5,000

Air Permeability (mD)

1,000
3

500

12
13

100

10

50

11

10

11A

5
4

1.0
0

10

20

30

40

50

60

70

80

90 100

% Connate water
1
2
3
4
5
6
7
8

=
=
=
=
=
=
=
=

Hawkins
Magnolia
Washington
Elk Basin
Tangely
Creole
Synthetic Alundum
Lake St John

10 =
11 =
11A =
12 =
13 =

Louisiana Gulf Coast


Miocene Age-Well A
Ditto-Wells Band C
North Belridge California
North Belridge California
Core Analysis Data
Dominguez Second Zone
Ohio Sandstone

Relative Permeability: A quantity (fraction) that describes the amount of impairment


to flow of one phase on another. It is defined in the two phase Darcy law (see notes);
depends on the Saturation of the phase; e.g. in two phase flow -> krw and kro depend
on Sw (since Sw + So = 1).
A schematic of the Two Phase Darcy Law showing the definition of Relative
Permeability is presented in Figure 8.
At steady-state flow conditions, the oil and water flow rates in and out, Qo and Qw,
are the same:

12

Figure 7
Correlation between (air)
permeability and the
connate water (Swc) for a
range of reservoir rocks

Glossary of Terms

At steady-state flow conditions, the oil and water flow rates in and out, Qo and Qw
are the same
Po
Pw

Qw
Qo

Qw
Qo

The two-phase Darcy Law is as follows:

Qw =

k.k rw .A . Pw
w L

k.k ro .A . Po
Qo =
o L

Figure 8
The two-phase Darcy Law
and relative permeability

Scematic of relative permeability,


krw and kro

1
kro

Rel.
Perm.
0

krw

Sw

The two-phase Darcy Law is as follows:


Where:

Qw and Qo = volumetric flow rates of water and oil


Where:
= cross-sectional area
Qw and QoAL = Volumetric
flow rates of water and oil;
= system length
A
= o Cross-sectional
w and
= system lengtharea;
k
= absolute
permeabilities
L
= System
length;
the pressure drops across the water and oil phases at
o =
w and o Pw and
= PWater
and
oil
viscosities;
steady-state flow conditions
k
= kro Absolute
permeabilities;
krw and
= the water
and oil relative permeabilities
Pw and Po = The pressure drops across the water and oil phases at
NB the Units for the
two-phase Darcy
are exactly the same as those in Figure 2
steady-state
flowLaw
conditions
krw and kro
= The water and oil relative permeabilities

Several further examples of relative permeabilities and capillary pressure are given
later in this glossary. Note that the Units for the two-phase Darcy Law are given in
Figure 2, Chapter 2.
Capillary Pressure: The difference in pressure between two (immiscible) phases;
defined as the non wetting phase pressure minus the wetting phase pressure; depends
on the saturation - for two phases capillary pressure, Pc(Sw) = Po- Pw (for a water wet
porous medium). The following figures show schematic figures for Capillary Pressure
(Pc(Sw)) and Relative Permeability (krw(Sw) and kro(Sw)) for a water wet system:

Institute of Petroleum Engineering, Heriot-Watt University

13

Capillary Pressure
Swc

Relative Permeability

Sor

Pc

Swc

Sor

krel
kro
krw

Sw

Sw

Mobility and Mobility Ratio: the mobility of a phase (e.g. w or o) is defined as


the effective permeability of that phase (e.g. kw = k . krw ; ko = k . kro) divided by the
viscosity of that phase;

k.k rw
k.k ro
w =
; o =

w
o

Mobility ratio, M, is given by:

M=

o k ro . w
=

w k rw . o

Fractional Flow: The Fractional Flow of a phase is the volumetric flow rate of
the phase under a given pressure gradient, in the presence of another phase. The
symbols for water and oil fractional flow are fw and fo and they depend on the phase
saturation, Sw:

fw =

Qw
Q
; fo = o ;;where Q T = Q o + Q w
QT
QT

The fractional flows play a central part in Buckley-Leverett (B-L) theory of linear
displacement which starts from the conservation equation:

Sw
f

= w ;
t
x

So
f

= o
t
x

Buckley-Leverett Theory: This mathematical theory of viscous dominated water


oil displacement is based on the fact that the velocity, vSw, of a fixed saturation
value Sw is given by:

f
vSw = v. w
Sw

14

Figure 9
Schematics of capillary
pressure and relative
permeability for a water-wet
system

Glossary of Terms

where v is the fluid velocity, v = Q/(A)) and (dfw/dSw) is the slope of the fractional
flow curve. The relationship between the fractional flow and Buckley Leverett theory
is illustrated in Figure 10.
Fractional Flow

Welge Tangent

Fractional
Flow
of Water,

"Buckley-Leverett" Saturation Profile

Sor

fw

Figure 10
Relationship between the
fractional flow function and
the Buckley-Leverett front
height

Sw

Flood Front
Height

Sor

Swf
Swc

Length, (x/L)
Swc

Swf

Water Saturation, Sw

1.5 Wettability and Fluid Displacement Processes

Wettability: This is a measure of the preference of the rock surface to be wetted by


a particular phase - aqueous or oleic - or some mixed or intermediate combination.
The Wettability of a porous medium determines the form of the relative permeabilities
and capillary pressure curves; a very complex subject which is still the subject of
much research. We refer to: Water wet, Oil wet and Intermediate wet systems in the
following definitions.
Water-Wet: Water-wet formation: Where water is the preferential wetting phase.
Water occupies the smaller pores and forms a film over all of the rock surface - even
in the pores containing oil. A Waterflood in such a system would be an imbibition
process (see below). Water would spontaneously imbibe (see below) into a waterwet core containing mobile oil at Sor, hence displacing the oil.
Oil-Wet: Oil-wet formation: Where oil is the preferential wetting phase. Oil occupies
the smaller pores and forms a film over all of the rock surface - even in the pores
containing water. A Waterflood in such a system would be a drainage process (see
below). Oil would spontaneously imbibe into an oil-wet core containing mobile
water at Swr, displacing the water.
Intermediate-Wet: An Intermediate wet formation is where some degree of
water wetness and oil wetness is shown by the same rock. Some different types of
intermediately wet system have been identified known as Mixed wet and Fractionally
wet. Both water and oil may spontaneously imbibe (see below) into such a system to
some degree and indeed this forms the basis for certain Wettability Tests known as
the Amott Test and the USBM Test (USBM => United States Bureau of Mines).
Drainage: A Drainage displacement process is when the non-wetting phase is
increasing. For example, in a water wet porous medium, drainage would be oil
displacing water. The drainage and imbibition capillary pressure curves and relative
permeabilities are different since these petrophysical functions depend on the saturation
history. A simple schematic of a drainage process is shown in Figure 11.
Institute of Petroleum Engineering, Heriot-Watt University

15

Qo
Qo
Oil Injection

Water Wet Core


at 100% Water
Oil Injection
Water Wet Core
Qo
at 100% Water

Figure 11
Drainage

Imbibition: An Imbibition displacement process is when the wetting phase is


increasing. For example,
in a waterWater
wet porous
medium, drainage would be water
Oil Injection
Wet Core
displacing oil as shown in Figure 12. at
The
drainage
100%
Waterand imbibition capillary pressure
curves and relative permeabilities are different since these petrophysical functions
depend on the saturation
Q history.
w

Water Injection
Q

Water Wet Core


at sor
Water Injection
Water Wet Core
at sor
Qw
w

Figure 12

Water Wet Core


Spontaneous Imbibition: This process occurs
at sorwhen a wetting phase invades a porous
Water Injection

Imbibtion

medium in the absence of any external driving force. The wetting fluid is sucked
in under the influence of the surface forces. For example, if we have a water wet
core at irreducible water saturation, Swr, then water may spontaneously imbibe and
displace the oil as shown in Figure 13.
Oil

The observed behaviour in a system of Intermediate Wettability is shown in Figure 14


Oil
where we
seeatthat
Core
s both phases can spontaneously imbibe under certain conditions.
wc

Core at swc
Core at swc

Water

Water imbibes into


core displacing
Water imbibes
oil-water
wet or into
core
displacing
intermediate wet
oil-water wet or
system
intermediate
wet
Water
imbibes
into
system
core displacing
oil-water wet or
intermediate wet
system

Oil

Water
Water

Core at swc

Figure 13
Spontaneous Imbibition

Core at sor

Oil

Water

Primary and Secondary Recovery Processes: Primary and Secondary processes


refer to the stage in the fluid displacement when one phase displaces another. For
example, in a water wet porous medium (--> means displaces) :

16

Drainage and Imbibition


Capillary Pressure

Drainage (d) and Imbibition (i)


Relative Permeabilities
Swc

Sor

Figure 14
Intermediate wettability.
Both water and oil may
spontaneously imbibe
into the core displacing
the other phase. Shows
both water wet and oil wet
character

Glossary of Terms

Primary drainage: is oil --> water from a core at 100% water saturation to Swr.
Secondary imbibition: is water --> oil from a core at Swr and mobile oil to Sor.
Examples: Figures 15 and 16 show schematics of typical Drainage and Imbibition
capillary pressure (Pc) and relative permeability (krw and kro) curves for a water wet
system. Primary Drainage (oil --> water from core at 100% water) and Secondary
Imbibition (water --> oil from core at Swr) processes are illustrated:
Drainage and Imbibition
Capillary Pressure

Drainage (d) and Imbibition (i)


Relative Permeabilities
Swc

Sor

krel

Pc

Drainage
kro

Figure 15
Drainage and imbibition
capillary pressures

Imbibition

krw

Sw

Sw

Figure 16
Drainage and imbibition
relative permeability char-

Relative Permeability, %

100
80
60

Drainage

40
Imbibition

20
0

20
40
60
80
100
Wetting Phase Saturation, %PV

acteristics

Examples: Further examples of experimental capillary pressures and relative


permeabilities in cores are shown for various processes (Drainage and Imbibition)
and wettability conditions (Water wet and Intermediate wet) in figure 17,18,19 and
20 on the following pages.

Institute of Petroleum Engineering, Heriot-Watt University

17

80
80
60
60

OilOil

Relative
Relative
Permeability,
Permeability,
Fraction
Fraction

100
100

40
40
20
20
0
00
0

Water
Water
20
40
60
80
20Water 40
60 %PV
80
Saturation,
Water Saturation, %PV

100
100

Figure 17
Typical water-oil relative
permeability characteristics,
strongly water-wet rock

1.0
1.0

0.01
0.01
WaWtearter

Relative
Relative
Permeability,
Permeability,
Fraction
Fraction

il il
O O

0.1
0.1

0.001
0.001

0.0001
0.0001 0
0

20
40
60
80
20 Saturation,
40
60 %PV
80
Water
Water Saturation, %PV

100
100

Figure 18
Typical water-oil relative
permeability characteristics,
strongly water-wet rock

80
60

r
Wa
te

40
20
0

18

Oil

Relative Permeability, Fraction

100

1.0

20
40
60
80
Water Saturation, %PV

100

Figure 19
Typical water-oil relative
permeability characteristics,
strongly oil-wet rock

il

Glossary of Terms

1.0

Wate
r

Relative Permeability, Fraction

il

Figure 20
Typical water-oil relative
permeability characteristics,
strongly oil-wet rock

0.1

0.01

0.001

0.0001

20
40
60
80
Water Saturation, %PV

100

Examples: Experimental Capillary Pressures in cores for various processes (Drainage


and Imbibition) and wettability conditions (Water wet, Oil Wet and Intermediate
Wet). Figure 21.

Institute of Petroleum Engineering, Heriot-Watt University

19

Water Wet
Venango core VL-2
k = 28.2 md

Capillary Pressure - Cm of Hg

40
32
24
16
2

8
0
0

40

Capillary Pressure Characteristics,


Strongly Water-Wet Rock.
Curve 1 - Drainage
Curve 2 - Imbibition

24
16

20

32

16

24

12
8
4
0

20
40
60
80
Water Saturation - %

100

Capillary Pressure Characteristics,


(After Ref. 30)

20

40
60
80 100
Oil Saturation - %

Oil-Water Capillary Pressure Characteristics,


Ten-Sleep Sandstone, Oil-Wet Rock
(After Ref. 29).
Curve 1 - Drainage
Curve 2 - Imbibition

Capillary Pressure - Cm of Hg

Capillary Pressure - Cm of Hg

32

0
0

100

20
40
60
80
Water Saturation - %

Oil Wet

48

Capillary Pressure - Cm of Hg

48

Intermediate Wet

16
1
8

0
-8

-16
-24

20
40
60
80
Water Saturation - %

100

Oil-Water Capillary Pressure Characteristics,


Intermediate Wettability.
Curve 1 - Drainage
Curve 2 - Spontaneous Imbibition
Curve 3 Forced Imbibition

20

Figure 21

Glossary of Terms

Examples: Relative Permeabilitites


Examples of experimental relative permeabilities in cores for Water Wet and Oil
Wet systems. Figure 22.
100
Relative Permeability, Fraction

40
20
0

Water

20
40
60
80
Water Saturation, %PV

60

40
20
0

100

Typical Water-Oil Relative Permeability


Characteristics,
Strongly Water-Wet Rock

Oil

60

80

Wa
te

80

Oil

Relative Permeability, Fraction

100

20
40
60
80
Water Saturation, %PV

100

Typical Water-Oil Relative Permeability


Characteristics,
Strongly Oil-Wet Rock

1.0

1.0

O
il

0.001

0.0001

Figure 22

0.1

Wate
r

Relative Permeability, Fraction

0.01
Water

Relative Permeability, Fraction

il
O

0.1

0.01

0.001

0.0001
0

20
40
60
80
Water Saturation, %PV

100

Typical Water-Oil Relative Permeability


Characteristics,
Strongly Water-Wet Rock

Institute of Petroleum Engineering, Heriot-Watt University

20
40
60
80
Water Saturation, %PV

100

Typical Water-Oil Relative Permeability


Characteristics,
Strongly Oil-Wet Rock

21

EXAMPLES: Relative Permeabilitites

A simple table summarising the typical characteristics of water-wet and oil-wet


relative permeabilities is given below.
WATER WET

OIL WET

Swc

mostly > 20%

< 15% (Often < 10%)

Sw where
krw = kro
(Points A on Figure 23)

Sw > 50%

Sw < 50%

krw at Sro
(Points B on Figure 23)

krw < 0.3

krw > 0.5 (approaching 1)

This is shown schematically in Figure 23:


100

100
Water-Wet
Reservoir

80

Oil-Wet
Reservoir

80

Relative Permeability,
% of Air Permeability

Oil

Oil

60
40

40

20

20

Swi

20

60

r
Wate

40
60
80
Water Saturation, %

100

A
Swi
0

Wa

20

te

40
60
80
Water Saturation, %

In Water-Wet System
Sw mostly > 20%

In Oil-Wet System
Sw < 15%

At Point A: kro = krw ; Sw > 50%


krw at Sor / kro at Swi < 30%

At Point A: kro = krw ; Sw < 50%


krw at Sor / kro at Swi > 50%

100

1.6 Oil Recovery Methods, Waterflood Patterns and Sweep Efficiency


Here we refer to the method used to develop the reservoir as follows:

Primary Depletion - allowing the reservoir to produce under the original


reservoir energy i.e. by natural expansion. If the pressure falls below the bubble
point (Pb), then gas appears and the primary depletion process is known as solution
gas drive:

Secondary Recovery - where reservoir pressure is supported by injection,


usually of water in waterflooding but early gas injection may be considered also as
secondary recovery In addition to supporting the pressure (maintaining reservoir
energy), water or gas injection also displaces oil directly:

22

Figure 23
Influence of wettability on
relative permeability (after
Fertl, OGJ, 22 May 1978)

Glossary of Terms

Tertiary Recovery or Enhanced Oil Recovery (EOR) or Improved Oil


Recovery (IOR) - this refers to a range of methods which are designed to recover
additional oil that would not be recovered by either Primary or Secondary
Recovery.
Such methods include:
Thermal Methods
Gas Injection
Chemical Methods
Microbial Methods

- steam, in-situ combustion,..


- N2, CO2, hydrocarbon gas injection (usually after a waterflood)
- surfactant, polymer, alkali,..
- using bugs to recover oil!

Waterflood Pattern: On-land Waterflooding is often carried out with the producers
and injectors in a particular pattern. This is known as pattern flooding and examples
of such patterns are: Five Spot, Nine Spot, Line Drive etc. as shown schematically
in Figure 24.
Injection Well
Production Well
Pattern Boundary

Regular Four-Spot

Five-Spot

Figure 24
Examples of areal patterns
of injectors and producers
(pattern flooding)

Normal Nine-Spot

Skewed Four-Spot

Seven-Spot

Direct Line Drive

Inverted Nine-Spot

Inverted Seven-Spot

Staggered Line Drive

Areal Sweep Efficiency: The Areal Sweep Efficiency refers to the fraction of areal
reservoir that is swept at a given pore volume throughput of displacing fluid as shown
schematically in Figure 25. For example, the Areal Sweep Efficiency at Breakthrough
for various processes (Waterflooding, Gas Displacement and Miscible flooding) is
shown as a function of mobility ratio in Figure 26:
Institute of Petroleum Engineering, Heriot-Watt University

23

High Areal Sweep

High Areal Sweep

Figure 25
Schematic of areal sweep
efficiency

Poor Areal Sweep

Poor Areal Sweep

Breakthrough Areal Sweep


Efficiency, %

100

Breakthrough Areal Sweep


Efficiency, %

100
90

90
80
70

80
70

60
50
0.1

60

WaterOil
GasOil
Miscible

50
0.1
WaterOil

1.0
Mobility Ratio

1.0
Mobility Ratio

Figure 26
Areal sweep efficiency at
breakthrough in a five
spot pattern for various
displacement processes

10.0

10.0

GasOil
Vertical Sweep Effi
ciency: The Vertical Sweep Efficiency refers to the fraction of
vertical section (orMiscible
cross-section) of reservoir that is swept at a given pore volume
throughput of displacing fluid. This is function of the heterogeneity of the system
(e.g. stratification), the fluid displacement process (e.g. waterflooding, gas injection)
and the balance of forces (e.g. importance of gravity). It is shown schematically in
Figure 27.

24

Glossary of Terms

Good Vertical Sweep

Figure 27
Schematic showing vertical
sweep efficiency

Poor Vertical Sweep


(By gravity over-ride or the presence
of a high-k streak in this case)

1.7 Terms Used in Numerical Reservoir Simulation

Mass Conservation: This is a general principle which is used in many areas of


computational fluid dynamics. It says that:
(mass flow rate into a block) - (mass flow rate out)
= (the rate of mass accumulation in that block)
A reservoir simulation model (for 1, 2 or 3 phases) is basically: A mass conservation
equation combined with Darcys law.
Black Oil Model: Different types of formulation of the transport equations for
multiphase/multicomponent flow are used to simulate the various recovery processes;
by far the most common is the Black Oil Model which can simulate primary
depletion and most secondary recovery processes. A black oil simulation model is
one of the most common approaches to modelling immiscible two and three phase
(o, w, g) flow processes in porous media; it treats the phases rather like components;
it does not model full compositional effects; instead, it allows the gas to dissolve
in the other two phases (described by Rso and Rsw); however, no oil is allowed to
enter the gas phase.
Grid Structure: This refers to the geometry of the grid being used in the numerical
simulation of the system. This grid may be Cartesian, radial or distorted and may be
1D, 2D or 3D (see notes).
Spatial Discretisation: This is the process of dividing the grid in space into divisions
of x, y and z. In reservoir simulation, we always chop up the reservoir into
blocks as shown in the gridded examples below and then we model the block
block flows.

Institute of Petroleum Engineering, Heriot-Watt University

25

Temporal Discretisation: This is the process of dividing up the time steps into
divisions of t.
2D Areal Grid: This is a 2D grid structure as shown in Figure 28 which is imposed
looking down onto the reservoir. For a Cartesian system, it would divide up the x
and y directions in the reservoir into increments of x and y.
y

W1

W2

W3

x
y
Just 1 x z-block in 2D Areal Grid

Figure 28
Perspective view of a
2D areal (x/y) reservoir
simulation grid: W = well

2D Cross-Sectional Model: This is a 2D grid structure which is imposed on a


vertical slice down through the reservoir. For a Cartesian system, it would divide up
the x and z directions in the reservoir into increments of x and z. Cross-sectional
calculations are carried out to asses the effects of vertical stratification in the system
and to generate pseudo-function for upscaling. (Figure 29).

Figure 29

3D Cartesian Grid
The 3D Cartesian Grid is the most commonly used grid when constructing a relatively
simple model of a reservoir or a setion of a reservoir. This is shown in Figure 30.

26

Glossary of Terms

Producer

Water Injector

Figure 30
A 3D Cartesian grid for
reservoir simulation

z
(Variable)

Transmissibility: The transmissibility between two grid blocks is a measure of how


easily fluids flow between them. The mathematical expression for two phase flow
between grid blocks i and (i+1) is (for water):
(i+1/2) Boundary

Block i
Sw i

Block i+1
Qw

krw i (kA)i

Figure 31

Sw i+1
krw i+1 (kA) i+1

k
(P Pi )
Q w = ( kA
kA )i +1/ 2 rw
. i +1
x
w Bw i +1/ 2
where the inter-grid block quantities are averages at the interfaces (where i+1/2 denotes
this block to block interface. The single phase Transmissibility,, Ti+1/2 , is given by:

Ti +1/ 2 =

( kA)i +1/ 2
x

and the full Water Transmissibility,, Tw,i+1/2, between the two grid blocks is given
by:

Tw, i +1/ 2 =

( kA)i +1/ 2 k rw
x

k
= Ti +1/ 2 . rw

w Bw i +1/ 2
w Bw i +1/ 2

Institute of Petroleum Engineering, Heriot-Watt University

27

Therefore, we may write:

Qw = Tw, i +1/ 2 ( Pi +1 Pi )
The water transmissibility is clearly made up of two parts each of which is an
average between the blocks. The single phase part is (k.A)av and the two phase part
is [krw/(w.Bw)]av
(k.A)av - a Harmonic Average between blocks is taken for the single phase part of
the transmissibility (see Chapter 4; Section 3.2).
[krw/(w.Bw)]av - this term is more complicated. For the average relative permeability
term, [krw]av an Upstream Weighting is used; For [(w.Bw)]av the Arithmetic Average
between blocks is taken. (See Chapter 4; Section 3.3).
Numerical Dispersion: The spreading of a flood front in a displacement process
such as waterflooding, which is due to numerical effects, is known as Numerical
Dispersion. It is due to both the spatial (x) and time (t) discretisation or truncation
error that arises from the gridding. This spreading of flood fronts tends to lead to early
breakthrough and other errors in recovery. How bad the error is depends on several
factors including the actual fluid recovery process being simulated e.g. waterflooding,
water-alternating-gas (WAG) etc. (See Chapter 4; Section 2.2).
Grid Orientation: The Grid Orientation problem arises when we have fluid flow
both oriented with the principal grid direction and diagonally across this grid as
shown schematically in Figure 32. Numerical results are different for each of the
fluid paths through the grid structure. This problem arises mainly due to the use of
5-point difference schemes (in 2D) in the Spatial Discretisation. It may be alleviated
by using more sophisticated numerical schemes such as 9-point schemes (in 2D).

I = Injector
P = Producer

Local Grid Refinement: Local Grid Refinement is when the simulation grid is
made fine in a region of the reservoir where (LGR) quantities (such as pressure or
saturations) are changing rapidly. The idea is to increase the accuracy of the simulation
in the region where it matters, rather than everywhere in the reservoir. E.g. LGR ==>
close to wells or in the flood pilot area but coarser grid in the aquifer.
28

Figure 32
Flow arrows show the fluid
paths in oriented grid and
diagonal flow leading to
grid orientation errors

Glossary of Terms

Hybrid Grid LGR: Hybrid Grids are mixed geometry combinations of grids which
are used to improve the modelling of flows in different regions. The most common
use of hybrid grids are Cartesian/Radial combinations where the radial grid is used
near a well. Hybrid Grid LGR can be used in a similar way to other LGR scheme.
Examples: A simple example of LGR and Hybrid Grid structure is shown in figure
33.
Injector

Coarse Grid
in Aquifer

Producer

Figure 33
Schematic of local grid
refinement (LGR)

Hybrid Grid

Distorted Grids: A Distorted Grid is a grid structure that is bent to more closely
follow the flow lines or the system geometry in a particular case.
Corner Point Geometry: In some simulators (e.g. Eclipse), the option exists to
enter the geometry of the vertices of the grid blocks. This allows the user to define
complex geometries which better match the system shape. This option is known as
Corner Point Geometry and it requires that the block block transmissibilities are
modified accordingly.
The idea of Corner Point Geometry is illustrated schematically in figure 34:

Institute of Petroleum Engineering, Heriot-Watt University

29

Corner Point Geometry


Coordinates of
Vertices ( ) Specified
Block ( ) centres

Highly Distorted Grid Blocks

Block <-> Block


Transmissibilty

Fault
L1
L2
L3

L1
L2
L3
L4

L4

Distorted Grid

30

Figure 34
Grid structures for Faults
and Distorted Grids

Glossary of Terms

1 Extented Refinement

2 Imbedded Refinement (Rectangular)

3 Variable Refinement (Radial)

4 Hybrid (Radial in Rectangular) Local Grid Refinement


Around Vertical Wells

5 Hybrid Local Grid Refinement (Horizontal Wells)

Figure 35
Types of local Grids
Institute of Petroleum Engineering, Heriot-Watt University

31

History Matching: History Matching in numerical simulation is the process of


adjusting the simulator input in such a way as to achieve a better fit to the actual
reservoir performance. Ideally, the changes in the simulation model should most
closely reflect change in the knowledge of the field geology e.g. the permeability
of a high perm streak, the presence of sealing faults etc. The observables which
are commonly matched are the field and individual well cumulative productions,
watercuts and pressures.
Examples: Examples of history matching of pressure and water-oil-ratio (WOR) in
two reservoirs are Figure 36. Note that in the WOR match the first pass was very
inaccurate but that eventually a suitable match was found. A vitally important point is
that a good history match must be obtained for the right reason. It may be possible
to get a satisfactory match for the wrong reason i.e. by adjusting a variable that is not
the primary cause of the mis-match (indeed, this is very often the case). However,
such a model will eventually have very poor predictive properties.

(a)

1.0

3600
3400

0.8
Final Match
Pressure (psi)

3000

Watercut

0.6

0.4

5
7
Time, Days x 10

2600
2400
2200
0

6 7
Year

9 10 11 12

(b)

Observed Data
1

2800

2000

First Trial

0.2

0.0
0

Calculated
Field Data

3200

3600
11

3400

Calculated
Field Data

3200
Pressure (psi)

3000

Initial and Final Matches of WOR of a Well in a


Highly Stratified Reservoir

2800
2600
2400
2200
2000
0

6 7
Year

9 10 11 12

(c)
3600
3400

Calculated
Field Data

3200
Pressure (psi)

3000
2800
2600
2400
2200
2000
0

6 7
Year

9 10 11 12

Final Pressure Matches of Typical Khursaniyah Field Wells:


(a) Reservoir AB
(b) Reservoir C
(c) Reservoir D (MP)

1.8 Numerical Solution of the Flow Equations in Reservoir Simulation

Finite Differences: When the derivative in a differential equation is approximated


as a difference equation as follows:
32

Figure 36
A field of a history match
of watercut and well
pressures; redraw from
Mattax and Dalton (1990)

Glossary of Terms

S (Si Si 1 )

x i
x
then this is referred to as a Finite Difference approximation. In this example, which
is illustrated below, (S/x)i is the derivative of Saturation (S) with respect to x at grid
point i ; Si and Si-1 are the discrete values of S at grid points i and i-1, respectively;
x is the size of the spatial grid.
Si-1

Slope = s
x

S(x,t)

Si
Si+1
x

xx ....

i-1

i+1

Figure 37

Linear Equations: When finite difference methods are applied to the differential
equations of reservoir simulation, a set of linear equations results. These have the
form:

A. x = b
where A is a matrix of coefficients, x is the vector of unknowns and b is the (known)
right hand side. Expanded up, this set of linear equations has the form:
a11 x1
a21 x1
a31 x1
a41 x1

+
+
+
+

a12 x2
a22 x2
a32 x2
a42 x2

+
+
+
+

a13 x3 ....
a23 x3 ....
a33 x3 ....
a43 x3 ....

............
an1 x1 + an2 x2 + an3 x3 ....

+ a1n xn =
+ a2n xn =
+ a3n xn =
+ a4n xn =

b1
b2
b3
b4

+ ann xn = bn

Direct Solution Of Linear Equations: A Direct Solution method is when the linear
equations are solved by an algorithm which has a fixed number of operations (given
N, the number of linear equations [unknowns]). If the equations have a solution,
then, in principle, a direct method will give the exact answer, x(true), to the machine
accuracy. E.g. Gaussian Elimination
Iterative Solution Of Linear Equations: An Iterative Solution method is when the
linear equations are solved by an algorithm which has a variable number of operations.
A first estimate of the solution vector x(0) is made and this is successively refined to
converge to the true solution. In a convergent iterative method, then x(v) x(true) as
v . It is because of this iterative process that a variable number of steps may

Institute of Petroleum Engineering, Heriot-Watt University

33

be required depending on how accurate the answer must be. Normally, the iterative
method would be carried out until:
| x(v) - x(true)| < Tol.
where Tol. is some small pre-specified tolerance. E.g. Line Successive Over-relaxation
(LSOR)
Grid Ordering Schemes: When the simulation grid blocks are ordered in various
ways, the structure of the non-zeros in the sparse matrix, A, is different. Advantage
can be taken of the precise structure when solving these equations. E.g. Schemes
known as D2 and D4 ordering.

1.9 Pseudo-Isation and Upscaling

Upscaling: The process of reproducing the results of a calculation which is carried out
on a fine grid on a coarser grid is known as Upscaling. The basic idea of upscaling
is shown schematically in Figure 38. The input properties at the coarser scale must
take into account the flow effects of the smaller scale structure. These coarser scale
properties then become pseudo-properties.

34

Glossary of Terms

Fine Grid Layered Model


High Sw

Low Sw
Oil

Upscaling or Pseudo-Isation

Coarse Grid Layered Model

Fine Grid
Coarse Grid

% OOIP

Oil Recovery

Figure 38
Basic idea of upscaling

Time

Pseudo-Property: This refers to the value of a property or function (e.g. permeability,


relative permeability..) which is an average or effective value at a certain scale usually the grid block scale. For example, we might put the value kx = 150 mD in a
simulator grid block which is 200 ft x 200 ft x 30 ft. Clearly, this incorporates a large
amount of geological substructure and permeability may vary very significantly in
different parts of this block.
Pseudo-Relative Permeability: This is probably the most important pseudo-property
that is used in reservoir simulation. It refers to the effective relative permeability
in the simulation model at the grid block scale and is shown schematically in
Figures 39 and 40. It must incorporate the effects of all the smaller scale geological
heterogeneity, the balance of forces (viscous/capillary/gravity) and certain numerical
effects (numerical dispersion). Methods for calculating the Pseudo-rel perm include:
Jacks et al, Kyte and Berry, Stone etc. Newer methods are based on tensor pseudorelative permeabilities (Pickup and Sorbie, 1994).
Institute of Petroleum Engineering, Heriot-Watt University

35

Geopseudos: When the fluid flow upscaling is performed in the correct context of
the sedimentary structure up from the lamina, laminaset, bedform.. scales, then the
approach is known as the Geopseudo Methodology. This has been developed in
work at Heriot-Watt which has extended more conventional approaches by putting
in the geology.
Figure 40 shows a simple example of a pseudo relative permeability showing holdup
of fluid.

Fine Grid Layered Model


High Sw

"Rock"
Relative Permeabilities

krel
Low Sw

Sw

Oil

Upscaling or Pseudo-Isation

Pseudo-Relative
Permeabilities

krel
Sw
Coarse Grid Layered Model

Fine Grid
Coarse Grid

% OOIP

Oil Recovery

Time

36

Figure 39
Basic idea of upscaling or
pseudo-isation

Glossary of Terms

Water Flow

Rel Perms.
Sor

Sw
Swc

Sw
Rel Perms. ?

No Water Flow
Figure 40
A simple example of a
pseudo relative permeability

krw

kro

Sor
Sw
Swc

1.10 Numerical Simulation of Flow in Fractured Systems

Fractured System: In this context, we imply systems (such as in many carbonate


reservoirs) where small scale conductive fractures occur but most of the oil is in the
rock matrix. In certain non-porous fractured rock reservoirs (e.g. fractured volcanics),
it is possible to have all the oil in the fractures but these are much less common.
Typically, in porous fractured systems: fracture porosity, f = 0.1 - 1% of bulk volume
(i.e. as a fraction f = 0.001 - 0.01).
Features of Fracture Geometry: The main geometric features of fractures which
are thought to affect fluid flow are the: fracture orientation, width, conductivity,
connectivity and spacing (or fracture density). The interface between the fracture
and the matrix will also play a very important role in multiphase flow and fluid
displacement processes.
Stylolites: Stylolites are frequently found in limestones. They are laterally extensive
features formed by grain-to-grain sutured contact caused by the phenomenon of
pressure solution. These features may significantly reduce vertical permeability thus
causing systems containing them to have very low (kv/kh) ratios (at certain scales).
Vugs: Vugs are dissolution holes in a carbonate rock caused by diagenetic
reactions.
Dual Porosity Models: These are the most widely used simulation models for
modelling flow in fractured systems. They have separate conservation/flow equations
for the matrix and the fractures and matrix fracture flow is represented by Transfer
Functions. They are most frequently used to model multiphase flow in fractured
carbonates.
Variants of this model allow for; (i) flow only in fractures and (ii) flow in both
fractures and matrix.
Discrete Models of Fracture Systems: A more recent approach to flow in fractured
systems tries to represent the fractures explicitly as oriented planes with various
shapes in 3D. Single phase (tracer) flow models of this type are used to model radioInstitute of Petroleum Engineering, Heriot-Watt University

37

nuclide transport in fractured media. However, multi-phase models of this type are
not commercially available at present.
Transfer Function: The function which describes the oil flow rate between the matrix
and the fractures is known as the Transfer Function. Approximate analytical equations
for this function have been suggested by Birk, Boxerman and Ahronovsky.
Sudation: When oil is recovered from the matrix blocks in a fracture by a combination
of gravity and capillary forces, the recovery mechanism is sometimes referred to as
Sudation.

1.11 Miscellaneous-Vertical Equilibrium, Miscible Displacement and


Dispersion

Vertical Equilibrium: The concept of Vertical Equilibrium (VE) is quite widely


used in reservoir engineering. It takes several forms, two of which are listed below
(and illustrated schematically in Figure 41:
A: the pressure gradients in a particular direction (x, say), P/x, are all equal locally
in a long system. See Figure 41a.
B: there is virtually instant crossflow vertically - nearly infinite - compared with the
horizontal flow. See Figure 41b.
The VE assumption is often made in order to simplify the mathematical analysis of
certain fluid flow problems in reservoir engineering.
Vertical Equilibrium (VE) is known to apply in long thin systems (where x >>
z, in Figure 41b). More accurately, the VE limit is approached as the scaling group,
RL ; where:

x k
R L = . z
z k x

1/ 2

and kz and kx are the vertical and horizontal permeabilities, respectively. In practice,
if RL is > 10, then VE is a very good assumption.

38

Glossary of Terms

A: Equality of Layer Pressure Gradients (Where x >> z)


Layer 1
Layer 2

Layer 3
x

P
x

P
x

P
x

B: Instantaneous/Infinite Vertical Crossflow

Figure 41
Schematic views of vertical
equilibrium

Miscible Displacement: Whereas oil and water are immiscible fluids (i.e. they
do not mix and are separated by an interface), some fluids are fully Miscible (i.e.
they mix freely in all proportions). When a gas (or other fluid) is injected into an oil
reservoir and the fluids are miscible, this is referred to as a Miscible Displacement.
When two fluids (e.g. gas and oil) are fully miscible (go = 0), the local pressure and
the pressure gradients are the same (there is no capillary pressure since there is no
interface).
The mixing between the solvent and the oil can occur locally by Dispersion and
by Fingering (see Viscous Fingering below). The displacement is described by a
generalised Convection-Dispersion Equation where the mixing viscosity, (c) is a
function of the concentration of the solvent, c (or oil). Often, the solvent viscosity
is below that of the oil (i.e. s < o) which tend to cause an instability to develop in
the displacement known as Viscous Fingering.
The Miscible Flow Equations: These comprise of a Pressure Equation and a Transport
Equation. The pressure equation is derived by inserting Darcys Law (with a viscosity
dependent on solvent concentration) into the continuity equation. The transport
equation is a generalised convection-dispersion (or convection-diffusion) equation.
Continuity equation:

.u = 0 (assuming incompressible flow)


Darcy Equation:

u=

k
.P
P
( c)

where u is the Darcy velocity, c is the miscible solvent concentration and k is the
Institute of Petroleum Engineering, Heriot-Watt University

39

permeability tensor.
The pressure equation is then given by:

.u = .
.
P
P = q
(c )

or
k

.
.
P
P = q
(c )

where q represent any source/sink terms


The transport equation is the generalised convection-dispersion (diffusion)
equation:

c
cc
= .(D.c) v.
t
where D is the dispersion tensor and v is the pore velocity ( v =

u
).

Dispersion and Dispersivity: Hydrodynamic dispersion in a porous medium at


the small (core) scale is a frontal spreading or mixing which is due to various flow
paths which the fluid can flow along at the pore scale. This mixing is a diffusive
process since the growth of the mixing zone, Lf , tends to grow in proportion to t .
In a tracer core flood experiment, the Dispersion Coefficient, D, may be measured
by fitting the effluent profile to an analytical solution of the Convection-Dispersion
Equation (see figure 42). Units of D (cm2/s - at lab scale).
Dispersive mixing behaviour can also be seen from the mixing effect of heterogeneities
at larger scales in a porous medium. D, has been found to depend linearly on velocity
through the relationship: D = . v; where is the Dispersivity and has dimensions
of length. Dispersion is actually a tensor in rather the same way that permeability
( k ) in its general form is a tensor.

40

Glossary of Terms

Dispersive
Mixing Zone
C(x,t1)

Lf

Ce 1

Effluent
Concentration
Ce(1,t)

C
0
Figure 42
Schematic illustration
of dispersion and the
convection-dispersion
equation for simple tracer
flow

1
time (pv)

In situ concentration profile at time, t1; C(x,t1)


2

Described by Convection Dispersion Equation: C = D C2 - v C


t
x
x
C = Dimensionless Concentration (C/Co)
D = Dispersion Coefficient
v = Pore Velocity (Q/A)

Viscous Fingering: When a high mobility (lower viscosity) fluid displaces a lower
mobility (higher viscosity) fluid, a type of instability may develop known as Viscous
Fingering. For such a displacement, the Mobility Ratio (see above) is high and the
process may be observed in either miscible or immiscible displacements, although it
occurs more readily in miscible systems. Results from an experiment are shown in
Figure 43 where the dark fluid is high viscosity and the light fluid is low viscosity.
Clearly, viscous fingering leads to a poorer sweep efficiency in such floods.

Figure 43
Experimental
demonstration of viscous
fingering

Institute of Petroleum Engineering, Heriot-Watt University

41

PAPER NO. 28902B

HERIOT WATT UNIVERSITY


DEPARTMENT OF PETROLEUM ENGINEERING
Examination for the Degree of
Meng in Petroleum Engineering
Reservoir Simulation
Monday 17th April 1995

09.30

- 12.30

( 70% of Total marks )

NOTES FOR CANDIDATES


(1)

Answer ALL the questions and try to confine your answer to the space provided on the
paper.

(2)

The amount of space and the relative mark for the question will give you some idea of the
detail that is required in your answer.

(3)

If you need more space in order to answer a question then continue on the back of the
same page indicating clearly (by PTO) that you have done so.

(4)

The total number of marks in this examination is 262; this will be rescaled to give an
appropriate weighted percentage for the exam. The marks are relative and, together with
the space available, should give an approximate guide to the level of detail required.

(5)

There is a compulsory 15 minute reading time on this paper during which you must not
write anything.

(6)

You will be allocated 3 hours to complete this paper.

Q1.

List two uses of numerical reservoir simulation for each of the following stages of field
development.

(i)

At the appraisal/early field development stage


1.
(2)
2.
(2)

(ii)

At a stage well beyond the maximum oil production in a large North Sea field:
1.
(2)
2.

(2)

Q2.

At any stage in a reservoir development by waterflooding, the engineer may use material
balance calculations and/or numerical reservoir simulation. Under which particular cir
cumstances would you use each of these approaches?
(i) Material Balance?

(3)

(ii) Reservoir simulation?

(3)

Q3. Given all the problems and inaccuracies, which are known to exist in the application of
reservoir simulation, why do engineers still use it?

(2)
2

Q4
Water
injector
Water
injector

Producer

High k

Continuous or
Discontinuous
Shales ???

Low k

OWC

2000 m

Vertical scale
100 m

Reservoir X is a light oil reservoir (35 API) being developed by waterflooding. The reservoir
comprises a high average permeability massive sand overlying lower average permeability
laminated sands. The thickness of each sand is approximately equal and there are shales at the
interface of these two sands. However, the operator is uncertain if these shales are continuous or
discontinuous.
The following questions refer to Figure 1 above.
(i)

What would the main differences be between the cases where the shales in the above
reservoir were continuous and where they were discontinuous?

(4)

(ii)

Say briefly how you would go about investigating this using reservoir simulation.

(4)
3

(iii)

Suppose a reservoir engineer took 10 vertical grid blocks (NZ=10) in a simulation model
of this system. Would you expect the local (kv.kh) values in each of the main reservoir
sands to be similar or different? Explain your answer briefly.

Similar/different
(1)

Explain

(4)
(iv)

If this reservoir well had a gas cap, then gas coning might be a problem.

What is gas coning? Draw a rough sketch.

(4)
How would you use reservoir simulation to investigate this problem?

(4)
4

(v)

In which two ways would the grid used to investigate gas coning be different from that
which was used in the full field waterflooding simulation?

1.
(2)
2.
(2)

Q5.

Two of the main numerical problems/errors that arise in reservoir simulators are due to
numerical dispersion and grid orientation.

Explain each of these terms briefly saying - what it means, its origin and how we might get
round it. Draw a simple hand drawn sketch illustrating each.
(i)

Numerical Dispersion: Sketch

(5)
(ii)

Numerical Dispersion - Explanation?

(5)
5

(iii)

Grid Orientation - sketch

(4)
(iv)

Grid orientation - Explanation?

(4)
Q6.

A very simple single phase pressure equation is given by Eq. 1 below.


2
P P
= 2
t x

(i)

Eq. 1

Write down how this equation is discretised in an explicit finite difference scheme - briefly
explain your notation.

(4)

(ii)

If an implicit finite difference scheme was used to solve Eq. 1, then a set of linear
equations would arise which could be solved using either a direct or an iterative linear
equation solution technique. Briefly explain each of the bold terms above:

Implicit finite difference scheme

(4)
Set of linear equations

(4)
Direct linear equation solution technique

(4)
Iterative linear equation solution technique

(4)
7

Q7.

Statement: The Equations of Two Phase Flow can be derived easily simply by using
Material Balance and Darcys Law.

Explain this statement with reference to two phase flow - you do not need to actually derive the
equations and, indeed, you may not use any equations other than Darcys law.

(8)

Q8.

(i)

Draw a schematic sketch of a single grid block of size Dx by Dy by Dz, showing the
porosity f, the oil and water saturations So and Sw (only 2 phases present).

(2)

(ii)

(a)

Using the sketch in part (i) above, derive expressions for (a) the volume of oil in the grid
block, Vo; (b) the mass of oil in the grid block, Mo, introducing the formation volume
factor, Bo.
Vo?

(3)
(b)

Mo?

(3)

(iii)

Write an expression for the oil flux, Jo, saying briefly what it is, any units it might be
expressed in and explaining any terms you introduce.

(6)
9

Q9.

(i)

Name three ways in which a Black Oil reservoir simulation differs from a
Compositional simulation model.

1.

2.

3.

(6)
(ii)

Draw a simple sketch of a single grid block showing what is meant by a component and a
phase.

(4)

(iii)

Using the notation CIJ to denote the mass composition of component I per unit volume of
phase J (dimensions of CIJ are mass/volume), derive an expression - based on the
quantities labelled in your sketch in (ii) above - for the mass of component I in the grid
block.

(6)
10

(iv)

Give one example of (a) where you would use a Black Oil model and (b) where you would
use a Compositional simulation model.

(a)

Use Black Oil model?

(3)
(b)

Use Compositional model?

(3)

Q10. Figure 2: The figure below shows the basic idea of upscaling or Pseudo-isation.

"Fine" Grid Cross-Sectional Model

OIL

"Rock"
Propts.

"Coarse" Grid Upscaled or Pseudo-ised Model

"Pseudo-"
Propts.

With reference to Figure 2 above, answer the following:

11

(i)

What is meant by Upscaling with reference to the modelling of say a waterflood.

(2)
(ii)

What is the difference between rock relative permeabilities and pseudo-relative


permeabilities?

(4)
(iii)

In order to perform upscaling in reservoir simulation, we need both an Upscaling


Methodology and Upscaling Mathematical Techniques. Explain very briefly the
meaning of the bold terms.

Methodology

(4)
Techniques?

(4)

12

Q11. (i)

Sketch (roughly) a semi variogram for each of the following permeability models:

(a)

a correlated random field with a range of 100m and a sill of 10,000 mD2; and

(b)

a laminated system where the laminae are of constant width of 1cm and where the high
permeability = 2D and the low permeability = 1D. Label your sketches clearly.

(a)

(6)
(b)

(6)
(ii)

What can you deduce about the standard deviation of the correlated random field in (i)(a)
above.

(3)

13

(iii)

Sketch the correlogram for case (i) above.

(5)

Q12. Figure 3 below shows the sketch of simple 3 layer model.

k = 0.5 D

2 cm

k = 2.0 D

5 cm

k = 1.0 D

1 cm

(i) Calculate the effective permeability of the above model in the x-direction; show your working.

(5)

14

(ii)

Calculate the effective permeability of the above model in the y-direction, show your
working.

(6)
(ii) Suppose we put a very find grid (say of size 0.1 cm x 0.1 cm) on the 3 layer model in Figure 3
above. If we jumbled up all the grid blocks randomly so that the new model had no discernable
structure, would the new effective permeability be: greater than that in (i) and (ii)?; less than that in (i)
and (ii)?; in between these values? Explain your answer.

(6)

15

Q13. In miscible flow in a random correlated field, explain how the mixing zone grows with
time in each of the following cases (illustrate your answers with simple sketches):
(i) dispersive flow

(4)
(ii) fingering flow

(4)
(iii) channelling flow

(4)
(iv) On the same diagram below, sketch the expected type of fractional flow curve f(c) vs c) you
would expect for each type of flow.

(6)
16

Q14. In the Kyte and Berry pseudo-isation (upscaling) method, describe briefly (using a
diagram) how numerical dispersion is taken into account (no detailed mathematics is
required).

(10)
17

Q15. (i)

List the main categories in the hierarchy of stratal sedimentary elements - give one
short sentence explaining each of these.

(8)
(iii)

Describe which of the above length scales of sedimentary heterogeneity are likely to have
most significance for the following reservoir flow phenomena:

* Reservoir pay-zone connectivity:

(3)
* Gravity slumping or water over-ride:

(3)
* Vertical sweep efficiency:

(3)
* Residual/Remaining oil saturation

(3)

18

Frequency

Q16. You have been given the following distribution of core-plug permeabilities in a particular
reservoir unit which includes a higher permeability a fluvial channel sand overlying a
lower permeability deltaic sand:

100 400 600 800


Permeability (md)

With reference to Figure 4 above: (a) explain the probable reason that the permeability
distribution has the above form; (b) sketch the sort of permeability models (laminar, bed and
formation scale) you might use for the flow simulation of this unit.
(a)

(3)
(b)

(continue on the back of this sheet if necessary)

(10)
19

Q17. (i) Draw a sketch of water displacement of oil across the laminae in a water-wet
laminated system at (a) low flow rate (capillary dominated) and (b) high flow rate
(viscous dominated); in this sketch show where the residual remaining oil is and
give a sentence or two of explanation.

(a)

(8)

(b)

(8)

20

(ii)

Is the effective water permeability at residual (i) remaining oil saturation (across the
laminae) higher in case (a) or (b) in part (i) above? Explain.

(6)
(iii)

What are the implications of the results in (i) and (ii) above for upscaling in reservoir
simulation?

(6)

21

yyyy
;;;;
yyyy
;;;;
yyyy
;;;;
yyyy
;;;;

Model Solutions to Examination

O
N
t
N
O
no
TI
o
E: RA
td
M
T
bu ion
A
n
N GIS SE:
at
io
in
ct
RE UR
E:
se
am
:
R
s
ex
d
hi
CO AR TU
e t l the ishe
E
A
et
N
pl unti
fin
G
m
is
l
Co sea

SI

.:

8 Pages

Date:
Subject:

Reservoir Simulation
INSTRUCTIONS TO CANDIDATES
No.

Mk.

1. Complete the sections above but do not seal until the examination is finished.
2. Insert in box on right the numbers of the questions attempted.
3. Start each question on a new page.
4. Rough working should be confined to left hand pages.
5. This book must be handed in entire with the top corner sealed.
6. Additional books must bear the name of the candidate, be sealed and be affixed to
the first book by means of a tag provided

PLEASE READ EXAMINATION REGULATIONS ON BACK COVER

Answer Notes
#

=>

indicates one of several possible answers which are equally acceptable.

[] =>

extra information good but not essential for full marks - may get bonus.

Model Solutions to Examination

Q1

(i)

1. To perform broad scooping calculations which examine different


development options e.g. waterflooding, gas flooding etc.

2. To extend initial material balance calculations by examining


some other spatial factor such as well-placement or aquifer effect.

(ii)
#

1. To assess additional field management options such as infill drilling,


pressure blowdown etc.

2. To take the improved history match model which can be developed


after same development twice and to use this to assess various IOR
strategies e.g. gas injection, WAG or chemical flooding.

Q2
(i)
#

Because of its inherent simplicity you would virtually always apply


single material balance to assess your field performance - to see if DP
decline tallies with estimated field size, sources of influx and
production.

(ii)
Would be used when a more complex development strategy requiring
spatial information is essential e.g. well placement, assessment of shale
effects, gravity segregation etc.

Q3
(#)

Because it is the only tool we have to tackle complex reservoir


development/flow problems which extends material balance. Clearly, it
is much better than simple material balance alone.

Q4
(i)
The shale continuity strongly affects the hi/lo permeability layer
vertical communication (both pressure and fluid flow). Thus, it will
affect the effective kv/kh (lower or zero for continuous shales) and
will strongly influence gravity slumping of water in a waterflood. In
the situation above with high k on top, some vertical communication will
help recovery.

(ii)
Set up a simple 2D cross sectional model with , say, 50 blocks in the x direction and 10 vertical grid blocks - 5 in each layer. Run waterflood
cases with and without shales - and some in-between cases with
transmissibility modifiers set beween Tz = 0.0 0.01 0.1 0.5 1.0.
Compare water saturation fronts and recoveries as fraction of pv
water throughput. Result will allow us to assess the effects of the
shales in the waterflood.

(iii)
Different

Model Solutions to Examination

The high perm massive sand would have a small scale kv/kh ~1 which
would result in a similar larger scale value. In the laminated sands, the
small scale (say core plug scale) would have a low kv/kh of say 0.1 to
0.01 and this would result in a correspondingly lower kv/kh at the grid
block scale.

(iv)

Well

Gas

Oil
and
Water

Gas

Oil
and
Water

= perforations
Gas Coning It is the drawdown of the highly mobile (low mg) gas into
the perforations. Pattern is shown here in figure. Causes high GOR
production at a level well above the solution gas value.

Set up a near-well r/z geometry fine grid - possibly 50 layers and


set reservoir near-well rock properties e.g. Layering, Tz modifiers,
Rel. perms. etc.

Perform simulations to look at issues such as effect of rate, vertical


communication, gas/oil/water Rel. perms. etc
Generally needs a fine Dr, Dz grid, often finer near the well where
most rapid changes of Sg and pressure with time occur.

(v)
1. The geometry would be different: r/z for coning and cartesian or
corner point for full field.

2. The fineness of the grid would be different. Very fine for nearwell; much coarser for full field.

(Dimensionality too 2D vs. 3D)

Model Solutions to Examination

Q5
(i)

(ii)
Numerical dispersion is the artificial spreading of saturation fronts
due to the numerical grid block structure in the simulation. It arises
because we take large grids to represent moving fronts. It can be
improved by refining the grid (globally or locally) or by using improved
numerical methods.

(iii)
Wells same distance
apart in Figs A and B
L
I

Fluid tends to
flow along (parallel)
to the grids

Fig A

Fig B
I = injector ;

P = producer

(iv)
The injected fluid tends to flow parallel with the grid from the
injector (I) to the producer (P) - see previous page. This means that
early breakthrough and poorer recoveries are seen in A then in B
above. i.e.
Fig B
Actual Recovery
Fig A
%00IP
Producer

Pv or Time

Q6
(i)

Model Solutions to Examination

(ii)
2 P
In this scheme the spatial term in Eq. 1 i.e.
would be specified at
x 2
the new time level n+1

A set of linear equations is the following type of equation (e.g.3x3system)

a11 X1 + a12 X2 + a13 X3 = b1


a21 X1 + a22 X2 + a23 X3 = b2
a31 X1 + a32 X2 + a33 X3 = b3

where X1, X2, X3 are unknown - the as are a matrix of known


coefficients and bs are a known right-hand side.

A direct solution method (e.g. Gaussian Elimination


Elimination) is an algorithm
with a fixed number of steps which will solve these linear equations
(under certain conditions). [Typically forward elimination is applied to
get an upper triangular A* matrix and back substitution is then easily
applied to get the X solution]

In contrast, an iterative technique starts with a first estimate of the


unknown vector X (0) where the (o) denotes 0th iteration: This is then
improved by some algorithm to a better and better solution of the
original linear equations i.e. X(1) X(2) X(V) until the method
converges e.g.

/ X(V+1) - X(V)/ < small number TOL. [Methods such as the Jacobi, LSOR,
etc. are examples of this].
Q7
In block (i, j), then material balance
can be applied for each phase (e.g.
oil and water) for 2-phase flow.
mo
mw

mo
i, j

Mass
Accumulation of

Amount that

Amount that

oil over time

flows in over

flows out over

Dt

Dt

Dt

But amount that flows in/out is given by the pressure differences


between blocks i.e.

A.k.kro ( So )
Qoil ( i -1, j ) ( i , j )
Po - Poi -1 j

mo

i - 1 ij
2

Thus the two phase Darcy Law supplies the relation for volumetric flow
rate and pressure in the grid block. These volumetric flows can be
converted to MASS flows (x by density) and then put into the material

10

Model Solutions to Examination

balance equation to obtain a conservation equation and in pressure


equation for oil and water.

\ Material Balance + Darcys Law => 2-phase Flow Equation.

Q8
(i)

(ii)

(a)

Vo = Dx Dy Dz f So

(b)

11

(iii)

Q9

1. The black oil model essentially treats a phase (o,w,g) as the basic
conserved unit or pseudo component

2. Compositional models are based more correctly on the conservation


of components (CH4, C10, H2O etc.) - the black oil model simply treats
gas dissolution in oil through Rso - gas solubility

3. The compositional models incorporate a full PVT description of the


oil whereas the black oil model relies on the simple Rso type treatment.

12

Model Solutions to Examination

(iii)

(iv)
#

(a) Waterflood calculations in a low GOR - say 30 API - oil reservoir


with pressure maintenance.

(b) CO2 injection in a - say 36 API - light oil system [Condensate


system - gas recycling etc]

Q10
(i)
Upscaling in a waterflood essentially means getting the correct
(effective) parameters (-e.g. rel. perm.) for the larger scale grid blocks
which will reproduce a correct fine grid model.

(ii)
Rock relative permeabilities are meant to be the intrinsic
representative properties of a representative piece of reservoir rock
at the small (i.e. core plug) scale.

Pseudo rel. perms. are effective properties at the larger (usually


gridblock) scale which incorporate other effects and artefacts (e.g.

13

numerical dispersion, heterogeneity etc..) in addition to the intrinsic


rock rel. perms.

(iii)
Methodology
This is a geologically consistent approach to the task of upscaling. i.e.
data collection, sedimentological framework,

[The function of the methodology is to get the geologically + fluid]


mechanically right answer.

Techniques?
These refer to the actual mathematical algorithm to go from a fine
grid coarse grid. E.g. Kyte and Berry, Stones method, two phase
tensors etc

[N.B. This just needs to reproduce the fine grid result - even if it is
WRONG - at the coarse scale]

14

Model Solutions to Examination

Q11

(i)

(ii)
It takes a lag distance of about the range to see the field variability
(standard dev. - i.e. ~ 100mD) of the field.

15

(iii)

lag

Q12
(i)

(ii)
The effective permeability is clearly the harmonic (thickness weighted) average as follows:

16

Model Solutions to Examination

(iii)
The keff in the randomised model would be between the two answers
in (i) and (ii) above (the answer in (ii) being the lower).

e.g. Strictly in a randomised distribution of permeability the average


value tends to the geometric average (kg) in 2D
kg - is less than the arithmetic (along layer) answer.
kg - is greater than the harmonic (across layer) answer.

Q13
(i)
Note - we take the same contour values (c= 0.1, 0.5, 0.8) in all sketches
below.

17

(ii)

(iii)

(iv)

18

Model Solutions to Examination

Q14 With no maths

Consider only ACTUAL flows of Qw and Qo across this interface


interface only. Thus, if the fine grid water (say) flows have not
reached the coarse grid interface

then Qw = 0. => set krw= 0

When there is oil or water flow e.g. water flow.

19

Q15
(i)
Lamina The simplest unit within which we can assume (almost)
homogeneous k. (length mm m)
Lamina set A collection of the above (cm m) e.g. core.

hi k
lo k

Bedform How lamina sets are joined together geometrically to


form 3D beds.

20

Model Solutions to Examination

Eroded/? top sets

e.g.

Tabular cross-bedding

2m

~50cm

Bottom sets

or climbing ripples

Para-sequence/sequence-stacks of bedforms

(ii)

Para sequence - sequence scale

At parasequence - sequence - also bed form influence

Para sequence - bed form

Lamina set - bed form

Q16
(a) There is a double peak - the bimodality probably arises from the
lower perm plugs from deltaic sands, and the higher permeability plugs
from the fluvial channel.

21

(b)

laminated sand
pseudo

Tightly laminated
deltaic sands
A
Crossbeddes
fluvial
channel
-stacked crossbeds

2 Scale pseudo-isation - inclined cross bed pseudo A


- bedform pseudo B

Q17
(i)

(a)

hi

lo

hi

lo

hi

lo

Slow Flow
CAPILLARY DOMINATED

Water flow direction

High water Sw in
LOW perms in a
water-wet system

Sw

HIGH "remaining"
oil in hi k
Spontaneous water inhibition into the LOW k laminae occurs in
Pc-dominated flow. This traps oil in the HIGH k laminae behind the
front where it is well above "residual" but it can't move because the
Rel. Perm. to oil in the low k water-filled laminae is so low.

22

Model Solutions to Examination

(b)

hi

lo

hi

lo

hi

lo

VISCOUS DOMINATED
WATERFRONT

"Fast" Flow
of water

Water flow direction

High water Sw
LOW perm in a
water-cut system

Sw

Note at Viscous dominated conditions a water front goes through


which reduces the oil in all layers to its local "residual" level.
Recovery of oil is better in this case since it is not "stranded" by
downstream capillary imbibition.

(ii)
It is higher in case (a) for the reasons already explained.
[I give a slightly over-detailed answer to part (a) and (b) above].

(iii)
The central implications are twofold.

(a) The two phase pseudo relative perms. are highly anisotropic for
such laminar systems. Along and across layer water displacement in
laminar system gives widely different pseudos.

23

Along

Across

(b) The levels of remaining oil can be vastly different in laminar


systems which, in simulation/upscaling, moves the pseudo rel. perm. end
points. (see above).

24

Model Solutions to Examination

25

Course:- 28117
Class:- 28912

HERIOT WATT UNIVERSITY


DEPARTMENT OF PETROLEUM ENGINEERING
Examination for the Degree of
Meng in Petroleum Engineering
Reservoir Simulation
Tuesday 20th April 1999
09.30 - 13.30
( 100% of Total marks )

NOTES FOR CANDIDATES


1.

This is a Closed Book Examination.

2.

15 minutes reading time is provided from 09.15 - 09.30.

3.

Examination Papers will be marked anonymously. See separate instructions for


completion of Script Book front covers and attachment of loose pages. Do not write your
name on any loose pages which are submitted as part of your answer.

4.

This Paper consists of 1 Section.

5.

Attempt all Questions.

6.

Marks for Questions and parts are indicated in brackets

7.

This Examination represents 70% of the Class assessment.

State clearly any assumptions used and intermediate calculations made in numerical
questions. No marks can be given for an incorrect answer if the method of calculation is
not presented. Be sure to allocate time appropriately.

Q.1: Consider the following statement which is made referring to a reservoir development plan
for a field which has been in production for some time:
A reservoir engineer should always apply Material Balance calculations and should usually but not always - use Numerical Reservoir Simulation.
(i) Why should you always perform some sort of material Material Balance calculations ?
..............................................................................................................................
..............................................................................................................................
..............................................................................................................................
...........................................................................................................

(4)

(ii) What is the main disadvantage of using material balance calculations in reservoir
development?
..............................................................................................................................
...........................................................................................................

(2)

(iii) In the above context, explain when you would use Reservoir Simulation and when you may
not use it. Give an example of each case.
When you would use Reservoir Simulation + Example:
..............................................................................................................................
..............................................................................................................................
..............................................................................................................................
...........................................................................................................

(4)

When you may not use Reservoir Simulation + Example:


..............................................................................................................................
..............................................................................................................................
..............................................................................................................................
...........................................................................................................

(4)

Q.2: Various types of 2D and 3D grids are used in reservoir simulation calculations. Describe
what you think the best type of grid is for performing calculations on each of the reservoir
processes described below and say why.
Reservoir processes:
(i) Modelling of likely gas and water coning and its effect on (vertical) well productivity in a
light oil reservoir with a gas cap and an underlying aquifer;
(ii) Simulating a large number of options in an injector/producer well pair in a gas injection
scheme where the objective is to look at the effects of formation heterogeneity on gas - oil
displacement and to develop some pseudo relative permeabilities to use in a full field model;
(iii) Carrying out an appraisal of an entire flank of a complex faulted field which has several
injector and producer wells.
(i) Which grid?............................................................................................................................
.................................................................................................................................................... (4)
Why?...........................................................................................................................................
.....................................................................................................................................................
.....................................................................................................................................................
..................................................................................................................................................... (4)

(ii) Which grid?............................................................................................................................


..................................................................................................................................................... (4)
Why?...........................................................................................................................................
.....................................................................................................................................................
.....................................................................................................................................................
..................................................................................................................................................... (4)
(iii) Which grid?..........................................................................................................................
.................................................................................................................................................... (4)
Why?..........................................................................................................................................
....................................................................................................................................................
....................................................................................................................................................
.................................................................................................................................................... (4)

Q.3: Numerical dispersion and grid orientation are two of the main numerical problems that
occur in reservoir simulation. Explain in your own words, with the help of a simple sketch,
the meaning of each of these terms:
(i) Numerical dispersion ? Sketch:

(5)
Numerical dispersion ? Explanation:
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (5)
(ii) Grid orientation ? Sketch:

(5)
Grid orientation ? Explanation:
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (5)
5

Q4.

Figure 1 below shows the results of a series of 6 waterflooding and gas flooding
calculations (labelled A - F) in a 2D vertical cross-sectional numerical simulation model.
Results are plotted as Oil Recovery at 1PV Injection vs. 1/NZ , where NZ is the number
of vertical grid blocks in the simulation. Assume (a) that the number of grid blocks in the
x-direction (NX) is sufficiently large and is constant in all calculations; and (b) that the
axes of the graph are quantitative.
Figure 1

2D cross-section
Key
gas injection
waterflood

No. vertical grid


blocks = NZ

Oil recovery
at 1PV
injection

60%
A
50%

D
E

40%
F
0.1

0.2

0.3

0.4

0.5

Inverse number of grid blocks (1/NZ) -->

Answer the following questions:


(i) How many vertical grid blocks (NZ) were used in case F? .........

(3)

(ii) Do the simulated waterflood and gas flooding calculations become more optimistic or
pessimistic as we take more vertical grid blocks in the calculation?
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (4)

Q4. (continued)
(iii) Extrapolate each of the calculations to NZ --> for both the waterflood and the gas flood
on Figure 1. Estimate the % recovery for each and the incremental recovery of the gas flood
compared with the waterflood. Comment on the implications of your result for carrying out a
gas flooding project in this reservoir.
Estimations...........................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (6)
Comment:
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (4)

End of Q.4
Q5. on next page

Q. 5

Figure 2 below shows a control volume - block i - for single-phase compressible flow in
1D. The quantities qi+1/2 and qi-1/2 are the volumetric flow rates of the fluid at the
boundaries of block i. All grid blocks are the same size in the x-direction (x) and the
cross-sectional area is constant, A = y.z (where y and z are the block sizes in the yand z-directions). The density and porosity are denoted by symbols - and respectively.
Figure 2

i-1
Area
=A
= y. z

i+1

q i-1/2

qi-1/2

x
x
With reference to the above figure,
(i) Write a clear mathematical expression for the change in mass in block i over a time step t
due to flow:
Change in mass due to flow over time step t =
......................................................................................................................................................
...................................................................................................................................................... (6)

(ii) Write a clear mathematical expression for the change in mass in block i from the beginning
of the time step to the end (i.e. the accumulation):
Difference in mass in block i over time step t =
......................................................................................................................................................
...................................................................................................................................................... (6)

(iii) Considering the above two expressions, what equation can you now write from material
balance ?
......................................................................................................................................................
...................................................................................................................................................... (6)

Q.5 (continued)
(iv) Are there any assumptions in the equation you have just written in part (iii) above? Briefly
explain.
......................................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (4)
(v) Now use the single-phase Darcy Law in the equation you wrote in (iii) above and show how by taking Limits x, t --> 0 - you obtain the pressure equation for single-phase compressible
flow: Show the steps you take.

(8)
(vi) If you have written down the answer to part (v) above correctly, then you should have
written down a non-linear partial differential equation (PDE). What does non-linear mean in
this context and explain in physical terms what the main problem is with this sort of equation.
Non-linear?...................................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (4)
The problem?...............................................................................................................................
......................................................................................................................................................
...................................................................................................................................................... (4)
9

Q.6: (i) From the four assumptions listed below, show clearly how Equation 1 arises from the
non-linear equation you derived in Q.5 part (v) above. Indicate clearly where you use each
of the assumptions in your derivation.
2
P P
=
2
t x

Equation 1

The term is a constant (Greek kappa - not permeability, k). Other quantities which may be
used are Cf - the fluid compressibility, - the density; k - the permeability; - the fluid viscosity;
- the porosity.
Assumptions:

1. The rock is incompressible.


2. Permeability (k) and viscosity () are constant.
3. The fluid has a constant compressibility, C f =

1
.
P

P
4. Pressure gradients are small - hence 0
x
2

(i) Answer:

(12)
(continue on the back of this page if necessary indicating that you have done so)
10

Q.6 (continued)
(ii) From your answer to part (i) above, write down what constant is in terms of the other
constants.
= .............................................................................................................................................. (4)
(iii) Using the notation in Figure 3 below, apply finite differences to Equation 1 above and define
the discretised spatial derivative at the new time level, (n+1) (i.e. an implicit method). Show each step
in your working and show clearly how this leads to a sparse set of linear equations.
Figure 3

P in+1

n+1

x
Time
level

All x and
t fixed.

P in

n
i-1

i+1

Space --->

(12)
(continue on the back of this page if necessary indicating that you have done so)
11

Q.6 (continued)
(iv) We can write down the set of linear equations that arises from applying finite differences to
the flow equations as follows:
A.x = b

Equation 2

where A is a matrix, x is the vector of unknowns and vector b is known. Write out Equation 2
explicitly for three equations and rearrange these to show how a simple iterative scheme can be
formulated. Say very briefly how this is solved. Give one advantage and one disadvantage of
an iterative scheme.

(continue on the back of this page, if necessary, indicating you have done so)
(10)

12

Advantage?......................................................................................

(2)

Disadvantage?.................................................................................

(2)

Q.7: The oil flux, Jo, into and out of a grid block is shown in Figure 4 below. Other quantities
are denoted as follows: Oil saturation So; porosity, ; Formation volume factor, Bo; Oil
density at standard conditions, osc; Darcy velocity of oil, vo (similar quantities apply to the
water phase).
Figure 4

Oil Flux
Jo

Jo
z

x
x

x + x

(i) Write an expression for the oil flux, Jo, giving any possible units.

(6)
(ii) The concentration of oil, Co, is defined as the mass of oil per unit volume of reservoir. Write
an expression for Co in terms of the quantities defined above.

(6)
13

Q.7 (continued)
(iii) Prove that, in 1D two-phase flow, then for the oil phase:
(Jo/x) = - (Co/t)

Equation 3

showing your working clearly.

(10)
(continue on the back of this page, if necessary, indicating you have done so)

14

Q.7 (continued)
(vi) State the two-phase Darcys law for oil using (Po/x) for the pressure gradient and kro for the
relative permeability, and substitute this into Equation 3 above and derive the oil conservation
equation.
Two-phase Darcy Law for the oil phase (in terms of Darcy velocity, vo)

(4)

Oil conservation equation:

(8)
End of Q.7

15

Q.8 In three-phase flow (oil, water and gas), we can define a gas flux, Jg, and a gas concentration,
Cg, in exactly the same way as was defined for oil in Q. 7 above.
(i) Explain physically why the gas flux and the gas concentration are more complicated than the
corresponding quantities for the flow of oil or water.
.........................................................................................................................................................
.........................................................................................................................................................
...................................................................................................................................................... (4)
(ii) Using Rso and R to denote the gas solubility factors, derive expressions for Jg and Cg showing
your working.

(12)
(continue on the back of this page, if necessary, indicating you have done so)
16

Q. 8 (continued)
(iii) Use your expressions for Jg and Cg in the conservation Equation 3 (for gas) to write down the
first step in obtaining the conservation equation for gas.

(6)
End of Q.8
Q.9: Explain what history matching is in a reservoir simulation of a field saying briefly how it is
done and what can go wrong.
History matching? How is it done? What can go wrong?
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................(10)
(continue on the back of this page, if necessary, indicating you have done so)
17

Q.10 (i) Write down the formulae for the arithmetic (ka), harmonic (kh) and geometric (kg) averages
of a permeability field with permeabilities k1, k2, .... kM. The number of data points you have = M.

ka =

kh =

kg =
(10)
(ii) State how you would use these averages for calculating the effective permeabilities in the
horizontal and vertical directions in the models in Figures 5(a) and 5(b) below.
Figure 5
(a)

(b)

For Figure 5(a): Horizontal keff & Vertical keff :

(8)
18

Q.10 (continued)
For Figure 5(b): Horizontal keff & Vertical keff :

(8)
(iii) Calculate the effective permeability for flow across the laminae in Figure 6 below. Show your
working.

20 mD, 1 cm

100 mD, 2 cm

10 mD, 1 cm

Figure 6

...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................(6)

End of Q.10

19

Q.11 The diagram in Figure 7 below shows a grid consisting of 2 coarse grid blocks, with 7 x 3 fine
blocks in each of these coarse blocks.
Figure 7

i=1 2

3 4

6 7

9 10 11 12 13 14

j= 1
2
z

3
x

Suppose you are calculating the pseudo relative permeabilities for the left coarse block using the
Kyte and Berry method. Assume that you have all the necessary information (saturations, pressures,
flows etc.) from a fine-scale (3x14) simulation.
(i) Show clearly on Figure 7 which part of the grid you would use for calculating the following
quantities and give a brief sentence of explanation:
the average water saturation

(3)

...........................................................................................................................................................
...........................................................................................................................................................(4)

the average pressure gradient

(3)

...........................................................................................................................................................
...........................................................................................................................................................(4)

the total flows of oil and water

(3)

...........................................................................................................................................................

...................................................................................................................................................... (4)
(ii) What is the formula for the average water saturation?
Sw =

20

(5)

Q. 11 (continued)

(iii) What is the weighting for the average pressure? Give a brief sentence of explanation.
(5)

..........................................................................................................................................................

........................................................................................................................................................ (3)

(iv) What is the formula for the total flow of water?

qw =

(5)

End of Q.11

21

Q.12 (i) By means of a simple sketch, show how a cubic packing of spherical grains is arranged.
Sketch:

(4)
(ii) Use the sketch to help you calculate the specific surface of the sample per unit volume of solid,
Ss (in m2/m3).
Specific surface - working:

Specific surface per unit volume of solid, Ss = ..................... m2/m3


(6)
(iii) If the grain radius is taken to be 10m, determine the porosity () of the sample

Porosity working:

Sample porosity () = .........


22

(6)

Q.12 (continued)
(iv) If the grain radius was 100m instead of 10m, what would the porosity () of the sample now
be?

Porosity = .................................................................................................................................... (3)

(v) What do the results of parts (iii) and (iv) above suggest concerning the porosity of cubic sphere
packs?

Ans.............................................................................................................................................

(2)

(vi) Write down the Carman-Kozeny equation in terms of the grain diameter (D), porosity (), and
the specific surface per unit volume of solid, Ss
Carman-Kozeny equation:

k=

(6)
(vii) Taking the grain radius to be 10 m and the tortuosity of the sample to be T=1, calculate an
approximate permeability in Darcies for a cubic packing of spheres (NB Use the porosity found in
Q.12 part (iii) in this calculation) .
Working:

Permeability = ................. Darcies. (1 Darcy = 1x10-12 m2)

(6)
23

Q.13 (i) Describe the meaning of the term contact angle and draw a rough sketch to illustrate
your answer

Contact angle?...................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
(4)
Sketch

(4)
(ii) If an oil/water meniscus is at equilibrium in a cylindrical capillary tube, what is the equation
that relates the capillary pressure, Pc, to the tube radius R and the contact angle ?

Pc=

(5)

What is this reduced form of the equation called?

..................................................................................................................................................... (3)

24

Q. 13 (continued)
(iii) Oil is introduced at the inlet face of the water-filled pore network shown in Figure 8 on the
next page. The numbers on Figure 8 refer to pore radii (in microns, m). The pores are taken to be
capillary tubes and are water-wet, the contact angle is = 60o in every pore and the oil/water
interfacial tension is 80 mN/m. The capillary pressure of the system is gradually increased and oil
begins to invade the network.
Shade in the pores that become oil-filled at each of the 4 capillary pressure values Pc1, Pc2, Pc3
and Pc4 in Figure 8 (NB 14.7 psi = 105 Newtons/m2).

Show any working out below:

25

Q. 13 (continued)

Figure 8: Shade in the pores that become oil-filled at each of the 4 capillary pressure values Pc1,
,
c
2
5
2
Pc3 and Pc4 below (NB 14.7 psi = 10 Newtons/m ).

Pc1= 0.6 psi

15
11

2
1

Oil

Water
10

20
3

3
12

12

Pc2= 1.2 psi

15
11

2
1

Oil

Water
10

20
3

3
12

12

Pc3= 3.0 psi

15
11

2
1

Oil

Water
10

20
3

3
12

12

Pc4= 10.0 psi

15
11

2
1

Oil

Water
10

20
3
12

3
12

(20)
26

Q.14 (i) Explain the differences between a drainage flood and an imbibition flood at the porescale, paying particular attention to the roles played by pore size, film-flow and accessibility to
the inlet (sketch each displacement in the boxes provided below to illustrate your answer).
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
..............................................................................

(10)

(continue on the back of this page, if necessary, indicating you have done so)

Imbibition
Sketch

(4)
Drainage
Sketch

(4)
27

Q. 14 (continued)
(ii) The strongly water-wet network in Figure 9 below is initially completely filled with oil and the
capillary pressure is so high that no water can currently imbibe. If the capillary pressure is slowly
decreased, so that water can invade via film-flow and snap-off, how is the residual oil distributed at
the end of imbibition (i.e. when Pc=0)? Shade in the residual oil using the network template
provided in Figure 10. The numbers on the bonds again refer to the tube radii in microns (m).
Note this network is deliberately different from that in Figure 8. Also, oil cannot enter the water
reservoir due to the presence of a water-wet membrane.
Figure 9
Initial distribution of oil

15
11

2
1

Water

Oil
10

20
3

3
12

12

Figure 10

Answer: Template for final oil distribution (shade in appropriately)

15
11

2
1

Water

Oil
10

20
3
12

3
12

(8)
End of Q.14

28

Q15. You have been supplied with the table of mercury injection data below.

Table 1
Mercury
saturation
0

P c (air/mercury)
(psia)
2

Oil saturation Pc(oil/water)


(psia)
0

0.2
0.4
0.5

4
5
6

0.2
0.4
0.5

0.6
0.8
0.9

7
20
60

0.6
0.8
0.9

(i) Sketch the air/mercury capillary pressure curve using the axes provided

(6)
(ii) Write down the equation used to re-scale mercury injection data to oil/water systems and use it
to complete Table 1 (assume the following values for interfacial tensions and contact angles:
mercury/air = 360x10-3 N/m, oil/water = 60x10-3 N/m, mercury/air = oil/water = 0o
Equation:

(6)
Now complete Table 1 at the top of this page

(8)
29

Q. 15 (continued)
(iii) Plot the oil/water capillary pressure curve using the axes provided below
Sketch.

(5)

(iv) Write down the equation that determines the Leverett J-function from capillary pressure datai.e. complete the following equation:

J(Sw) = Pc(Sw) x ..........................................

(Q. 15 is continued on the next page)

30

(5)

Q. 15 (continued)
(v)
Using the capillary pressure data shown below in Table 2, calculate an appropriate J-function
and complete the table below : assume the values k = 100 mD, = 0.1, interfacial tension, =10
x10-3 N/m, and contact angle, = 0o. Choose any suitable units but label your sketch clearly.
Sketch the J-function below.

Table 2
Complete the table
Wa te r S a tu ra tio n P c (p s ia )
1
0
0.8
2
0.6
4
0.5
6
0.4
10
0.2
20

J (S w)

(6)
Write down the form of the J-function here:

J(Sw)=
(4)

Sketch the J-function here:

(6)
31

Q.16 (i) Name the two most popular tests used to determine the wettability of a reservoir rock.
The .................................................................. Test

(2)

The .................................................................. Test

(2)

(ii) Give the three rules of thumb that can often be used to distinguish between water-wet and
oil-wet relative permeability curves.
Rule 1 ................................................................................................................................................
.......................................................................................................................................................(4)

Rule 2 ................................................................................................................................................
.......................................................................................................................................................(4)

Rule 3 ................................................................................................................................................
.......................................................................................................................................................(4)
(iii) Capillary pressure curves in Figure 11 below have been measured on two different core samples.
The two plots are shown below. Use your knowledge of wettability variations at the pore scale to
answer the following:
Figure 11
Pc

Pc

Sw

(A)

Sw

(B)

(iv) Which sample is probably the more water-wet? Explain your choice.
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
.......................................................................................................................................................(6)
32

Q.16 (continued)
(v) Give a physical explanation for the negative leg shown in the capillary pressure curve in Figure 11(A).
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
...........................................................................................................................................................
(6)

(End of examination paper)

Total number of marks possible = 462

33

Potrebbero piacerti anche