You are on page 1of 57

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

You might also like