Sie sind auf Seite 1von 340

G10RS

Reservoir Flow Simulation

Heriot-Watt University

Edinburgh EH14 4AS, United Kingdom


2

Produced by Heriot-Watt University, 2017

Copyright © 2017 Heriot-Watt University

All rights reserved. No part of this publication may be reproduced,


stored in a retrieval system or transmitted in any form or by any means
without express permission from the publisher.

This material is prepared to support the degree programmes in


Chemical and Petroleum Engineering.

Distributed by Heriot-Watt University


Reservoir Flow Simulation

©HERIOT-W ATT UNIVERSITY December 2017 v2


3

Acknowledgements
Thanks are due to the members of Heriot-Watt, School of Energy Geoscience
Infrastructure and Society who planned and generated this material.

We would like to acknowledge the assistance and contributions from colleagues


across the University and students in preparing this and support material.

©HERIOT-W ATT UNIVERSITY December 2017 v2


4

Overall Contents

Topic 1: Introduction to Reservoir Simulation

Topic 2: From Material Balance to Reservoir Simulation

Topic 3: Gridding in Reservoir Simulation

Topic 4: Derivation of the Flow Equations

Topic 5: Solution of the Flow Equations

Topic 6: Modelling and Upscaling

Topic 7: Petrophysical and Fluid Data for Reservoir Simulation

Topic 8: History Matching

Topic 9: Simulation of Reservoir Recovery Processes

©HERIOT-W ATT UNIVERSITY December 2017 v2


1

Topic 1 – Introduction to Reservoir


Simulation
Contents
1. Introduction ............................................................................................................ 3
1.1 What is a simulation model? ................................................................................ 3
1.2 What is a Reservoir Simulation Model? ............................................................... 5
2. The Task of Reservoir Simulation ........................................................................... 8
2.1 Steps Required to Set up a Reservoir Simulation ............................................... 10
2.2 Output from a Reservoir Simulator .................................................................... 12
2.2.1 Example output from a Reservoir Simulation ............................................. 13
2.3 Synthetic Example .............................................................................................. 14
3. Reservoir Simulation at Appraisal and in Mature Fields .......................................... 16
3.1 Appraisal Stage ................................................................................................... 16
3.2 Mature Field Development ................................................................................ 17
3.3 Late Field Development ...................................................................................... 18
4. Case Studies .......................................................................................................... 18
4.1 Middle Eastern Carbonate Reservoir (Sibley et al, 1997) ................................... 18
4.2 Schiehallion Field Study (Govan et al, 2006) ...................................................... 20
4.3 Cerro Fortunoso Field Study (Thompson et al, 2014) ........................................ 23
4.4 Conclusions from Case Studies ........................................................................... 27
5. Simulation Software ............................................................................................. 28
6. References ............................................................................................................ 28

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 Define a simulation model.

 Explain the difference between an analytical model and a numerical model.

 Outline the steps required to set up a reservoir simulation.

 Discuss the issues involved in choosing the complexity of a model.

 Contrast the use of simulation at different stages of a reservoir.

 Describe examples of reservoir simulation studies.

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
Reservoir simulation is used routinely in the oil and gas industry to help engineers to
plan how they should develop a reservoir, to predict how much oil (or gas) they may
recover, and also to investigate the effects of uncertainty in their models.

1.1 What is a simulation model?


Before learning about reservoir simulation in particular, we shall consider what a
simulation is, in general, and we shall study a simple example.

Definition: A simulation model is one which shows the main features of a real system,
but is simple enough to make calculations on.

The calculations may be analytical or numerical. Analytical calculations involve


mathematical functions (e.g. exp, sin, x2, etc), while numerical calculations involve
simplifying the problem, breaking it up into stages (time steps, grid blocks) and solving
it using an iterative approach.

Simple Example
This example has nothing to do with reservoir simulation, but is useful to illustrate the
two modelling methods. Imagine a colony of bacteria, where the growth of the colony
depends on the number of bacteria present, i.e.

dN
N , (1)
dt

where N is the number of bacteria at time, t, and  is a constant. This is a simple


differential equation, and the analytical solution is:

N  N0 exp  t  , (2)

where N0 is the number of bacteria at time t = 0. (If you are uncertain about this result,
try differentiating Equation (2)).

Figure 1 shows an example solution. Note, however, that as time increases the number
of bacteria gets infinitely large, according to this model, which is unrealistic. In the
case an actual colony of bacteria the growth rate would fall eventually because of, for
example, a lack of food. This demonstrates the limitations of our model. In fact,
models tend to be useful over certain conditions, but can fail outside their range of

©HERIOT-WATT UNIVERSITY December 2017 v2


4

applicability. There is a well-known saying “All models are wrong, but some models
are useful”!

Figure 1
The number of bacteria as a function of time – an example of an analytical
calculation. (See Tutorial 1.1.)

In the example above, the analytical solution is very simple. However, in more complex
problems (e.g. reservoir simulation), the model may be too complex to solve using an
analytical method. Instead, we use a numerical method. We assume that the number
of bacteria at a certain time is known, and we calculate the number a certain time later.
Assume that we have time steps 1, 2, … n, n+1, each separated by a time interval t.
Use the notation:

Nn = the number of bacteria at time step n,


Nn+1 = number of bacteria at time step n+1.

Then we discretise Equation (1).

N n 1  N n
Nn. (3)
t

This gives us a simple algorithm to calculate Nn+1 in terms of Nn,  and t.

©HERIOT-WATT UNIVERSITY December 2017 v2


5

N n1  N n 1  t  . (4)

We start at t = 0, with N = N0, and then iteratively calculate N at later times.

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. Tutorial 1 gives
you an opportunity to compare analytical and numerical solutions of the model of the
colony of bacteria.

1.2 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.

First consider the nature of reservoir rocks. Figure 2 shows a schematic diagram of the
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.

From these two figures, it can been seen that reservoirs can be highly complex. Figure
2 shows a wide variety of facies at scales occur over a scale of around 10 km. These
will have different poro-perm properties, which will affect the flow of fluid through the
reservoir. On the other hand, Figure 3 shows that rocks are very heterogeneous over
scales of mm – cm. Note that the air permeabilities (kair) range from 1mD to almost
3000 mD. Obviously reservoir models must be hugely simplified compared with a real
reservoir system.

Since reservoirs are complex, we use numerical simulation models. Spatially, the
model is split into grid cells, each of which has a set of properties – e.g. porosity, net-
to-gross and permeability. See Figure 4. It is important to realise that, although the
grid cells may be 100 m in length and perhaps 1 m in thickness, the properties in each
grid cell are constant. A reservoir simulation then solves equations (described in Topics
4 and 5) to estimate how the pressures and saturations change as the fluids are
injected into and produced from the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


6

Figure 2
A schematic diagram for the depositional system for the mid-Jurassic Linnhe and
Beryl formations.

©HERIOT-WATT UNIVERSITY December 2017 v2


7

Figure 3
Section of core from the Beryl Formation in the North Sea.

Figure 4
Example of a reservoir model, showing grid cells and the permeability distribution (in
mD). (MRST SAIGUP example; Lie, 2016.)

©HERIOT-WATT UNIVERSITY December 2017 v2


8

Figure 4 shows an example of a reservoir model. The grid follows the reservoir
architecture, and includes faults (discontinuities between grid cells). The colours of
the cells represent permeability. This is a synthetic, but realistic model which was used
to test the effect of geological uncertainty in oil reservoirs, in a project called SAIGUP
(Manzocchi et al, 2008). The model is included in the examples in the MRST software
package. The model has 40 x 120 x 20 grid cells. They are all 75 m x 75 m in the x- and
y- directions, and typically about 10 m in the z-direction, but vary depending on the
geological structure.

2. The Task of Reservoir Simulation


Apart from constructing a reservoir model and populating the cells with petrophysical
properties, we also need information on the fluid properties (PVT data), the relative
permeabilities and capillary pressure (if we have more than one phase flowing). In
addition, we need to know where the wells are located in the model, how they are
completed and how they are controlled (e.g. by rate or by bottom-hole pressure).
However, not all simulations models need to be complex – it depends on the question
you are trying to answer. Often engineers use simpler models to help them understand
a process better.

i. The simplest type of model is a tank model (Figure 5), which has just one grid cell
for the whole reservoir, and we are only concerned with the gross fluid flow into
or out of the reservoir. It is assumed in this case that the pressure equilibrates
rapidly across the reservoir. This is essentially a material balance model, which we
will discuss in Topic 2. The advantage of this model is that it is very simple. It can
address questions relating to the average field pressure, given the fluid fluxes into
or out of the reservoir. However, this model cannot simulate different pressures
in different parts of the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


9

Wells

Average Pressure = P
Average Saturations = So, Sw and Sg

Aquifer
Figure 5
A tank model.

ii. Slightly more complex than a model with a single tank, is a sector model with
several tanks. Material balance is applied to each tank. This sort of model is
frequently used when a reservoir is compartmentalised, and engineers are trying
to understand which compartments of a reservoir are in contact. Figure 6 shows
an example of a material balance model similar to one of the models used for the
Schiehallion Field (Dobbyn and Marsh, 2001). This field is discussed further in
Section 4.2. Again this model has the advantage of being simple and quick to run,
but it can show more than the single tank simulation.

Tank

Injector

Producer

Figure 6
Model with several connecting tanks.

iii. Thirdly there is a conventional reservoir simulation model, which may be 1D, 2D,
or 3D. Full field reservoir simulation uses a 3D model, as shown in Figure 4. The
advantage of a fine-scale reservoir simulation model is that you can model

©HERIOT-WATT UNIVERSITY December 2017 v2


10

heterogeneous poro-perm distributions, and the simulation can be used to address


more complex issues, such as deciding on well locations, or estimating time to
water breakthrough. On the other hand there is much work involved in setting up
a full-field simulation model.

It is important to realise that there is no single “right” way to model a reservoir. The
simplicity/complexity of the simulation should relate to the simplicity/complexity of
the question. Another important factor is the amount of data available. There is no
point in making a complex model if there is no data!

2.1 Steps Required to Set up a Reservoir Simulation


The following gives a very brief summary of the workflow used to set up a reservoir
simulation.

seismic well logs core data

fluid data

SCAL

well
locations
oil recovery

time

Figure 7
Data required for reservoir modelling and simulation.

a) Geological Model
This stage is normally the job of a geologist, rather than an engineer. Seismic
data is required to define the structure of a geological model: the horizons
between layers of rock are input. Also, fault information (locations and
throws) may be included.

b) Petrophysical Data

©HERIOT-WATT UNIVERSITY December 2017 v2


11

Well log data is analysed to determine porosities, net-to-gross, water


saturation and sometimes permeability at the well. In addition, core plug data,
if available, provides additional data on porosity and permeability.

c) Geological modelling
At this stage the model is gridded. The geologist, in consultation with an
engineer, should decide what the size of the grid cells should be. This is usually
a compromise between the geologist wanting a very fine grid to represent the
geological heterogeneity, and the engineer wanting to limit the number of grid
cells so that simulation runs don’t take too long. Then properties must be
assigned to the grid cells. Usually the rock type is modelled first (e.g. clean
sandstone, muddy sandstone, etc). Since there is little known about
reservoirs, apart from at wells, the rock types and their properties are usually
generated randomly, conditioning the values at the wells. (i.e. The values in
grid cells containing the wells are set to the average well log values.) After the
rock types are assigned, the petrophysical properties are generated at the
wells (net-to-gross, porosity and permeability). This data is also assigned
stochastically, constrained to well data.

d) PVT properties
Fluid samples from RFT data should be analysed to determine the PVT
properties. In compositional simulation, which we discuss in Topic 9, the
composition of the fluids and the equation of state (EOS) is required. However,
most simulations use a “Black Oil” model, in which the components of oil and
gas are not simulated. A typical PVT table for oil will contain Rs (dissolved gas
ratio), Bo (formation volume factor) and viscosity as a function of pressure. In
addition to inputting PVT tables, the reservoir simulator needs the surface
densities of the fluids.

e) SCAL data
If more than one fluid is present – e.g. oil and water – relative permeability
data and capillary pressure data is required. Topic 7 is devoted to describing
these data.

f) Initialisation of simulation
The reservoir simulator needs to be given some starting information: the initial
pressures and saturations. It is usually assumed that the fluids in the reservoir
start off in hydrostatic equilibrium, i.e. pressure increases with gz, where  is
density, g is the acceleration due to gravity and z is depth. The engineer needs
to put in the pressure at a datum depth, and also the fluid contacts (OWC and
GOC).

©HERIOT-WATT UNIVERSITY December 2017 v2


12

g) Well locations and completions


The grid cells containing the well head should be specified and also the cells
where the wells are completed. This links the simulation model to the surface
where fluids injected and produced.

h) Schedule for well control


The simulator needs to know how a well is controlled. This is either specified
as a rate (e.g. stb/day), or a bottom-hole pressure (BHP in psia). The period of
time for carrying out the simulation should also be specified. During the course
of a simulation, wells may be switched on or off to replicate a real scenario.

Figure 8
Example of a reservoir model. The grid is distorted (corner point geometry) with two
faults, and has over 25,000 cells (not very many, by today’s standards). Red is oil and
blue is water. There is a horizontal production well in the crest and two vertical
water injectors on the flank. Vertical projections of the x-z and y-z planes are shown
at the side.

2.2 Output from a Reservoir Simulator


There is a whole range of results which can be output from a reservoir simulator.
This is not an exhaustive list.
 Average field pressure as a function of time
 Field cumulative oil, water and gas production profiles with time
 Field daily (weekly, monthly, annual) production rates of each phase (oil, water
and gas)

©HERIOT-WATT UNIVERSITY December 2017 v2


13

 Water and gas injection rates (for individual wells or a whole field)
 Individual well pressures (bottom hole, or top hole using lift curves) with time
 Individual well cumulative and daily flow rates of oil, water and gas with time
 Either full field or individual well watercut, or gas-oil ratios with time
 Spatial distribution of oil, water and gas saturations throughout the reservoir as
functions of time.
 Spatial distribution of pressure as a function of time.

Some of the outputs can be compared with data from the field (if available) – e.g. well
rates, field rates, or well BHPs. (See Topic 8 on History Matching.) It is also very useful
to examine the saturation and pressure distributions as this helps to understand the
flow processes.

2.2.1 Example output from a Reservoir Simulation


The three diagrams below (Figure 9) show some examples of output from a reservoir
simulation. They are all taken from a paper by Sibley et al (1997). In this case the
geological models were “history matched” – i.e. parameters in the geological model
were adjusted so that the simulation results agree with the observations. More details
on history matching are given in Topic 8. (A further discussion of this field is given in
Section 4.1.)

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Figure 9
Example of some typical reservoir simulator output, from SPE 36540, Sibley et al,
1997, “Reservoir Modeling and Simulation of a Middle Eastern Carbonate”. Top:
history match for cumulative field oil and water recovery; Middle: history match for
average pressure; Bottom: history match for RFT data.

2.3 Synthetic Example


A simple example of how reservoir simulation might be used is illustrated in Figure
10. This is an imaginary case where the reservoir study is to consider the best of 4
options in Field A:
Option 1 – continue as present with the waterflood;
Option 2 – upgrade injection wells;
Option 3 – upgrade the injector and drill 3 new injector wells;
Option 4 – drill 2 new infill production wells.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

It is obviously much cheaper to model these four cases than to actually drill wells. The
important quantities are the oil recovery profiles for each case compared with the
scenario where we just proceed with the current development strategy (Option 1). Of
course, we do not know if our model is “correct”, so our predictions may not be
accurate. In fact, an important aspect of reservoir simulation is to assess each of the
various uncertainties associated with our model. This should lead to a range of
prediction profiles.

Figure 10 a) shows the reservoir model with the current wells – 3 producers and 3
injectors. Figure 10 b) shows the various options to be assessed. Reservoir simulations
are then performed for the 4 cases to determine the cumulative volume of oil
produced in each case. The results are shown in Figure 10 c), and indicate that the
option which produces the most oil is option 3, which was to upgrade the injectors and
drill 3 new injector wells. However, we must also assess the cost of each option.
Option 3 which involves drilling 3 wells is going to be more expensive than option 4
which involved only drilling 2 production wells. Therefore we may find that Option 4
has the greatest NPV (Figure 10 d). In other words, the aim of reservoir management
is always to maximise economic benefit.
new producers
a) b)

new injectors

c) d)

Figure 10
Synthetic Example. a) model with initial wells; b) model showing new wells; c)
cumulative oil with time; d) NPV for each option.

©HERIOT-WATT UNIVERSITY December 2017 v2


16

3. Reservoir Simulation at Appraisal and in Mature Fields


Reservoir simulation may be applied at all stages of development of a field. However,
simulation is used in different ways at different stages. We contrast reservoir
simulation at the appraisal stage and at a mature stage.

3.1 Appraisal Stage


At this stage reservoir simulation is used as a tool 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 facilities required to develop the field – e.g. a platform, subsea
development tied back to an existing platform, or a FPSO (floating production
storage and offloading – for an offshore field).
 The nature and capacities of plant sub-facilities, such as compressors for injection,
oil/water/gas separation capability.
 The number, locations and types of wells (vertical, slanted or horizontal) to be
drilled in the field.
 The sequencing of the well drilling program and the topside facilities.

During the appraisal stage the biggest, i.e. most expensive investment decisions are
made, e.g. the type of platform and facilities. Therefore, it is the most helpful time to
have accurate forward prediction of the reservoir performance. But at this time we
have the least amount of data. There is very little or no field performance history.
(There may be some extended well tests, though.) So it seems that reservoir
simulation has a built-in weakness: just when it can be most useful, it has the least data
to work on, and hence the predictions will be uncertain. However, even during
appraisal reservoir simulation can take us forward with the best current view of the
reservoir that we have at that time. Since the model is very uncertain at this stage,
usually a range of models is tested to give a range of predictions, as shown in Figure
11.

©HERIOT-WATT UNIVERSITY December 2017 v2


17

"Optimistic"
Cumulative Oil case
Recovery
(STB) Most probable
case

"Pessimistic"
case

2005 2010 2015


Time (Year)
Figure 11
An example of a range of predictions.

The sort of scenarios to be tested are:


 Different assumptions about the STOIIP – stock tank oil initially in place.
 Different values of reservoir parameters such as permeability, porosity, net-to-
gross, the effect of an aquifer, etc.
 Major changes in the structural geology or sedimentology of the reservoir – e.g.
sealing vs. “leaky” faults, the presence or absence of major fluvial channels, and
distribution of shales in the reservoir, etc.

3.2 Mature Field Development


We assume here that the field is in its “mid-life”, i.e. it has been in production for
several years (2 – 20 years), but it still has a reasonably long life ahead (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. This process can
be done on a continually updated basis. The main difference between this stage and
the appraisal stage 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. (History
matching is covered in detail in Topic 8.) The initial reservoir simulation model may
have been found to be “wrong”, in that it failed to predict some aspects of reservoir
performance – e.g. it may have failed to predict water breakthrough. (Often injected
water arrives at the oil producers before it is expected.) But that does not mean that
we should not have carried out a reservoir simulation in the first place. Creating a
model is a useful exercise, and helps us to understand the issues, and the accuracy of
a model can be improved gradually as more data becomes available.

At this stage in the development, typical simulation activities are:

©HERIOT-WATT UNIVERSITY December 2017 v2


18

 Carrying out a history match using the (now available) field production history in
order to obtain a better tuned model to use for future predictions of performance.
 Using the history match to re-visit the field development strategy in terms of
changing the development plan – e.g. infill drilling, adding extra water injection
wells, changing from water injection to gas injection, or using some method of
Enhanced Oil Recovery (EOR).
 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, 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. we may find that to match the history, we must
reduce the vertical flows (by lowering the vertical transmissibility). In this case, the
effect of gravity will be of less importance. Keith Coats refers to this as the
“educational value of simulation models”, and it is part of good reservoir
management that the engineer has a good grasp of the important reservoir physics
of their asset.

3.3 Late Field Development


We assume here that the field has only a few years left before abandonment. Firstly,
an engineer needs to determine whether the field is of sufficient economic importance
to merit a simulation study at this stage. However, there are two reasons why we may
want to launch a simulation study late in the field’s lifetime. Firstly, we may be able to
develop a strategy to give the field “a new lease of life” to keep it going economically
for a few years – e.g. by drilling more wells, well stimulation (to remove mineral scale)
or an EOR technique. Secondly, the cost of abandonment may be so high – e.g. if an
offshore structure has to be removed – that almost anything we do to extend the field
life and avoid this expense will be “economic”.

4. Case Studies
The following case studies have been chosen to help you to understand how reservoir
simulation fits into reservoir modelling and helps to improve reservoir understanding,
so that a field may be developed in the optimum way.

4.1 Middle Eastern Carbonate Reservoir (Sibley et al, 1997)


SPE paper 36540 (Sibley et al, 1997) describes the study of a cretaceous limestone
reservoir in the Arabian Peninsula. (The name of the reservoir is not disclosed.) The

©HERIOT-WATT UNIVERSITY December 2017 v2


19

field had been produced for many years (since 1957) and production was declining.
This study was carried out to assess future options for the field. Previously, a model
had been made by geologists and then passed to engineers. However, they could not
match the history. Therefore a new, integrated study was instigated, i.e. a
multidisciplinary team consisting of geoscientists and engineers worked together
throughout the project.

An important feature of this field is a diagenetic barrier. This is a layer of calcite-


cemented rock, which was formed at a previous oil-water contact. It has very low
permeability and so prevents water from the underlying aquifer providing support for
the producers. As a result, there was a large pressure decline in the upper producing
layers of the reservoir. Also, the reservoir consisted of layers due to cyclic deposition
of the carbonates. Figure 12 shows a schematic cross-section of the reservoir.

A huge amount of data was available for this field (although not all of the data was of
good quality). For example, there was slabbed core, thin sections, core analyses,
seismic, isotope analysis, open hole logs, TDT data (thermal decay time), RFT data
(repeat formation tests), field pressure data and oil and water production data. This
data helped geoscientists identify “flow units” – i.e. bodies of rock with similar flow
properties. The 3D geological model was built in terms of these flow units, rather than
depositional units. This was mainly because the diagenetic barrier cut across different
depositional units.

Figure 12
Schematic cross-section of the reservoir (Sibley et al, 1997).

©HERIOT-WATT UNIVERSITY December 2017 v2


20

Some examples of the history matching for this field were shown earlier, in Figure 9. It
can be seen that, thanks to the thorough understanding of the reservoir gained by the
data analysis, the history matches were good. The conclusion from this study was that
the integrated approach was essential for the understanding of the field. As a result
of the study, further development of the reservoir was planned. This included drilling
many new in-fill production wells, and drilling injector wells on the flank of the field
(Figure 13).

existing well
recently drilled well
new well
injector
C convert to injector

Figure 13
Structure map showing the locations of the old and new wells (Sibley et al, 1997). The
contours are at 50 ft intervals.

4.2 Schiehallion Field Study (Govan et al, 2006)


We consider the development of the Schiehallion field. This field is situated to the
North of the UK, West of the Shetland Isles (Figure 14). Information on this field may
be found in Govan et al (2006), SPE 96610. The field was discovered in 1993 and
development started in 1998. It is in a very hostile environment – the water depth is
400-500 m, and there are frequent storms. The geological setting is deep water
turbidite, so there are channel sands embedded in a lower permeability background
(Figure 15). There are also some E-W running faults which split the reservoir into
segments.

©HERIOT-WATT UNIVERSITY December 2017 v2


21

Shetland Isles

North Sea

Figure 14
Map showing the British Isles, with Shetland marked.

Figure 15
Schematic diagram of a turbidite system (Govan et al, 2006).

Since the channel sands have high permeability (~ 600 mD), it was initially thought that
there would be good recovery from this field. However, soon after the start of
production, the engineers realised that there were connectivity issues – the pressure
reduced faster than expected in the producers and there was an increase in gas-oil
ratio (GOR). The engineers suspected that the faults in the field were acting as barriers
to flow, and that they needed to get a better understanding of the field, and so needed
to collect more data. They performed extended well tests to determine the extent of
the reservoir compartments, and they shot repeat seismic surveys (4D seismic) to
monitor changes in the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


22

Dobbyn and Marsh (2001) described how a set of simple tank models were created to
investigate the connectivity between wells. These models were essentially a set of
material balance models (there was no flow within the tanks). This demonstrates that,
even though computers are currently capable handling models with millions of grid
cells, it is sometimes best to make simple models. Figure 16 gives an example of a tank
model. Compartments of the reservoir were represented by tanks, which usually
contained a well. Fluid was able to flow between tanks, and the transmissibility of the
connections was varied to match the history (i.e. observed pressure and production
profiles). Figure 17 shows the results of a history match to the pressure.

key

T## sand bodies


injectors
producers
T31a oil
gas

Figure 16
Example of a tank model (Dobbyn and Marsh, 2001).

Figure 17
Example of a history match to the pressure (Dobbyn and Marsh, 2001). The pink line
shows the results (in psia) of the material balance model, and the symbols show the
measured data. Also shown is the oil production and water injection in stb/day.

©HERIOT-WATT UNIVERSITY December 2017 v2


23

The advantages of running the tank models (material balance models) were:
 They helped engineers to understand how much water from one injector provided
support for a producer.
 They were easy to set up and fast to run. (There were only a few cells to consider.)
 The models were very useful in the early stages of field development – they helped
with the history matching of the simulation model.

The disadvantages of the material balance models were:


 They were less suitable for the later stages of development, when
o there were more wells,
o the fluid flow was more complex,
o there was communication over larger distances and longer time scales.
 Some tanks were not connected in the short term but were connected over
longer periods of time, once pressure differences had built up.

With additional data and a greater understanding of the connectivity in the reservoir,
the engineers were able to build a better reservoir simulation model (Govan et al,
2006). The aim of the detailed model was to assess the value of the water injectors,
by predicting how many barrels of oil could be produced per barrel of water injected.
This varied from 0.02 to 0.5, with an average of 0.17.

Govan et al (2006) also discuss the use of various different types of data used to
improve their model – 3D seismic data, 4D seismic data (i.e. repeat surveys), tracers
and pressure data. (See ppt file on the Schiehallion case study.)

4.3 Cerro Fortunoso Field Study (Thompson et al, 2014)


A study of the Cerro Fortunoso Field if described by Thompson et al (2014) in SPE paper
169369. It is a large, very heterogeneous field in Argentina. The depositional
environment is fluvial and the reservoir consists of thin, channelized sandstone beds.
The aim of the Cerro Fortunoso study was to assess whether waterflooding would be
economical throughout the remainder of the field.

The title of the SPE paper specifies that this is a field where normal modelling rules do
not apply. However, this statements refers to the problem of noise in the seismic data
and the fact that the reservoir is structurally complex. The SPE paper is still useful as
an example of a reservoir simulation study for this course. Note that this paper was
reviewed in JPT July, 2016 (Wilson, 2016).

©HERIOT-WATT UNIVERSITY December 2017 v2


24

The field has been on production since 1984 and had 240 production wells at the time
of writing the paper (2014). Most of the field was produced by primary depletion,
apart from a small area where water flooding was carried out, mainly for water
disposal. (Since this field is in a desert area, there is no cheaply available water supply,
so a careful study is required to evaluate the economics of water flooding.) The water
cut in the production wells was high though, because of an active aquifer on the flanks
of the reservoir. There is also a gas cap on the crest, which is 95% CO2.

Figure 18 shows some diagrams of the field. The left figure shows the well locations
with the contours representing depth. (Note that the reservoir is mainly above sea
level, though.) The centre image shows the different compartments. The one labelled
“NE” was used for the waterflood pilot and the modelling study. Finally some E-W
cross-sections are shown on the right. The colours illustrate gas, oil and water. Figure
19 focuses in on the NE sector. Note that the reservoir has very steep flanks and that
there is a large range of depths (700 m).

Figure 18
From left to right: the structure of the reservoir (red being highest and blue lowest);
the reservoir compartments; cross-sections through the reservoir showing gas (red),
oil (green) and water (blue).

©HERIOT-WATT UNIVERSITY December 2017 v2


25

Figure 19
Left: whole reservoir showing the location of the NE segment. Right: NE segment
plus wells. Note the big difference in height within the reservoir.

The main issue which had to be addressed in this field was the connectivity of the sand
bodies. RFT data showed that the pressures in some units were variable due to the
fact that the pressure had been depleted in some sandstones. They could not correlate
sands using just static data. From the static model, it looks like the sands in D and C
are disconnected. However, D was changed to an injector well and there was a
watercut response in well C. As they learnt more about the connectivity, they had to
update the static model grid and then re-propagate the facies model and the
simulation model.

Since many of the sandstone beds were very thin, they had to use thin grid cells, with
an average thickness of 1.25m. Because they needed to have several grid cells
between wells, the cells had a horizontal length of 30 m. This meant that there were
too many cells to simulate the whole reservoir, so only the NE sector was simulated.

SCAL data was available, but the relative permeabilities proved unreliable. They
produced water breakthrough where none was observed, so they increased the critical
saturation (i.e. the saturation below which there is no water flow). They also had
capillary pressure data. This was converted into an average J-function (Topic 7), which
proved to be more reliable for assigning the initial water saturation than the well log
data. This is because the thin sand beds were not properly resolved in the well logs,
which gave misleading results for the resistivity and therefore for the saturation.

©HERIOT-WATT UNIVERSITY December 2017 v2


26

History matching was carried out in several stages, firstly adjusting the size of sand
bodies, and then adjusting relative permeabilities. Figure 20 shows the history match
for the oil and water production rate in the NE sector. History matching is non-unique,
which means that a good match may be obtained from several different models. As
seen in Figure 20, the simulation overestimates the water production. This could be
due to a number of factors, such as: an underestimate of the connected hydrocarbon
pore volume, or an overestimate in the end-point water relative permeability.
Production Rate

Time
Figure 20
History match of oil and water rates in the NE sector. Note that the simulation was
controlled by oil rate, so the simulated and historical curves overlap.

Once a model has been history-matched adequately, forward predictions are run. In
this case they compared continuing with the existing flood with extending the water
injection. Figure 21 shows the saturations predicted for 2038. The oil saturation is
reduced in the case with the extended water flood.

The conclusions were that in order to understand the field, they have to integrate static
and dynamic data. Thanks to the modelling study, engineers were given the go-ahead
to expand the water flooding in this field.

©HERIOT-WATT UNIVERSITY December 2017 v2


27

Figure 21
Comparison of the predicted saturation distribution in 2038 (red = high oil saturation).
Left: continuing the existing waterflood. Right: extending the water injection.

4.4 Conclusions from Case Studies


These case studies demonstrate that to acquire a thorough understanding of a
reservoir, you need to integrate a lot of data. Firstly, there is geological data – seismic,
well logs, core measurements, etc. In addition for multiphase flow simulation,
information regarding fluid properties and petrophysical data (relative permeabilities
and capillary pressures) is also required. At the start of a field life, there may be little
data on which to base a model, so simple models are advisable (e.g. as in the
Schiehallion Field – Dobbyn and Marsh, 2001). However, later in field life more data is
available to give a better idea of the field structure and connectivity. In this case, a
detailed model and full reservoir simulation may be constructed (e.g. Govan et al 2006,
Thompson et al, 2014), in order to plan how the reservoir should be managed in the
future.

In all of these papers a large section is devoted to the collection and interpretation of
the data (geophysical, geological and petrophysical). You may think (as some
geologists do!) that reservoir simulation is merely a matter of pressing a few buttons
in a simulation package to predict the production. However, there is no guarantee that
you will be able to get reliable results from simulation. The role of a reservoir
simulation expert is:
 to determine the most appropriate type of model to run (tank model, complex 3D
model);
 to collaborate with geoscientists to decide on issues such as the grid resolution;
 to ensure that the simulation is adequate from a numerical point of view;
 to understand the limitations of modelling and simulation.

In order to do this, a reservoir simulation expert needs to understand:

©HERIOT-WATT UNIVERSITY December 2017 v2


28

 basic reservoir engineering


 the physics of fluid flow;
 how fluid flow equations are discretised to input into a reservoir simulator;
 how the equations are solved in the computer.

During this course you will be gradually introduced to these issues. You will not
become an expert within a few weeks, but hopefully this course will whet your appetite
to learn more.

5. Simulation Software
There are several commercial software packages available. Two commonly used
packages are:
 Eclipse
o Part of the Schlumberger suite for modelling and simulation
o Probably the best known simulation package
o Can be used for a wide range of problems
o Eclipse 100 is for “black oil” simulation (ignoring the composition of oil and gas)
o Eclipse 300 was developed for compositional simulation
 CMG Software
o Computer Modeling Group Ltd, Canada
o Includes IMEX for black oil simulation and GEM for compositional simulation
o Can also be used for a wide range of problems, including some which Eclipse
does not do (e.g. geochemical simulation).

It is usually very expensive to buy a license for these packages. In this course, you will
be using MRST, which stands for Matlab Reservoir Simulation Toolbox, which can be
downloaded free. The advantage of using MRST is that the code is not a “black box”,
like the codes in commercial software such as Eclipse. In MRST, you can view all the
codes, and you write Matlab scripts to call the routines. This gives more insight into
the process of reservoir simulation.

6. References
Dobbyn, A. and Marsh, M., 2001. “Material Balance: A Powerful Tool for
Understanding the Early Performance of the Schiehallion Field”, SPE 71819, presented
at the Offshore Europe Conference, Aberdeen, Scotland, 4-7 September, 2001.

©HERIOT-WATT UNIVERSITY December 2017 v2


29

Govan, A., Primmer, T., Douglas, C., Moodie, N., Davies, M. and Nieuland, F., 2006.
"Reservoir Management in a Deepwater Subsea Field - The Schiehallion Experience",
SPE 96610, SPE Reservoir Evaluation and Engineering, 9 (4), 382-290.

Lie, K.-A., 2016. “An Introduction to Reservoir Simulation Using MATLAB. User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, SINTEF ICT, www.sintef.no/mrst.

Manzocchi et al, 2008. “Sensitivity of the impact of geological uncertainty on


production from faulted and unfaulted shallow-marine oil reservoirs: objectives and
methods”, Petroleum Geoscience, 14 (1), 3 – 15.

Sibley, M.J., Bent, J.V. and Davis, D.W., 1997. “Reservoir Modeling and Simulation of a
Middle Eastern Carbonate Reservoir”, SPE 36540, SPE Reservoir Engineering, May
1997, 75-81.

Thompson, A., Garriz, A., Manestar, G., Vocaturo, G., Conoli, V. and Giampaoli, P., 2014.
“Intergrated Modeling of a Highly Structural and Complex Reservoir where the Normal
Modeling Rules Do Not Apply, Cerro Fortunoso”, SPE-169369, presented at the SPE
Latin American and Caribbean Petroleum Engineering Conference, Maracaibo,
Venezuela, 21-23 May, 2014.

Wilson, A., 2016.” Modeling of a Complex Reservoir Where the Normal Modeling Rules
Do Not Apply”, Journal of Petroleum Technology, July, 2016.

©HERIOT-WATT UNIVERSITY December 2017 v2


30

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 2 – From Material Balance to


Reservoir Simulation
Contents
1. Introduction ............................................................................................................ 3
1.1 Review of Basic Reservoir Engineering ................................................................. 3
2. Material Balance..................................................................................................... 5
2.1 Introduction to Material Balance ......................................................................... 5
2.2 Derivation of Simplified Material Balance Equations ........................................... 7
2.3 Conditions for the Validity of Material Balance ................................................. 11
3. Single Phase Darcy’s Law ...................................................................................... 12
3.1 The Basic Darcy Experiment ............................................................................... 12
3.2 Darcy’s Law with Gravity .................................................................................... 16
3.3 Radial Darcy’s Law .............................................................................................. 17
4. Darcy’s Law for Two-Phase Flow .......................................................................... 19
5. Fractional Flow Theory ......................................................................................... 21
5.1 Introduction to Fractional Flow .......................................................................... 21
5.2 Derivation of the Continuity or Mass Balance Equation .................................... 24
5.3 Buckley-Leverett Solution................................................................................... 25
5.4 Welge’s Method ................................................................................................. 27
6. Closing Remarks.................................................................................................... 29
7. References ............................................................................................................ 30
8. Appendix ............................................................................................................... 30
A1 Full Derivation of the Simplified Material Balance Equation .............................. 30
A2 Mathematical Note on Gradient and Divergence ............................................... 33
A3 Darcy’s Law in 3D ................................................................................................ 36

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:
 Define general reservoir engineering quantities, such as Bo and Rso
o This is revision from the reservoir engineering courses

 Calculate the STOIIP using a simple material balance model


o However, you do not have to memorise the material balance equation

 Describe the data required for material balance; discuss conditions for the material
balance equation to be valid; and present examples where material balance
cannot be used

 State Darcy’s Law for single phase flow


o Including the effect of gravity

 Derive Darcy’s Law for radial flow

 State Darcy’s Law for two-phase flow

 Derive the continuity equation for two immiscible, incompressible fluids

 Derive the Buckley-Leverett Equation and use Welge’s Method to calculate the
shock front height

 Calculate the time to breakthrough and the recovery efficiency at breakthrough


o You do not have to memorise these two equations, though

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
This Topic reviews some basic concepts of reservoir engineering that must be familiar
to the simulation engineer, which you have covered already in the Reservoir
Engineering courses. We start with Material Balance (G19RB, Topic 5) and the
definition of the quantities which are necessary to carry out such calculations: , co, cf,
Bo, Swi etc. You will perform simple material balance calculations in Tutorial 2.1 for this
Topic. Then the respective roles of Material Balance and Reservoir Simulations are
compared.

Then we go on to consider basic engineering associated with fluid flow: the single-
phase Darcy’s Law (relating flow rate and pressure gradient), including the effect of
gravity and also considering radial flow. We next look at two-phase flow, and review
a very useful method for estimating two-phase immiscible flow in a simple 1D model,
namely fractional flow and Buckley-Leverett Theory. Tutorial 2 for this Topic is on
Buckley-Levertt theory.

1.1 Review of Basic Reservoir Engineering


Although you are familiar with the basic concepts of reservoir engineering from
G19RA and G19RB, it is useful to include a review here:
a) it makes the course more self-contained;
b) it allows emphasis on the complementary nature of “conventional” reservoir
engineering and reservoir simulation;
c) we would like to review some of the flow concepts (Darcy’s Law, etc) in a manner
useful for
i. a study of Buckley-Leverett fractional flow theory,
ii. the derivation of the flow equations later in this course (Topic 4).

You need to be familiar with the following properties:


 density – mass/unit volume (lb/ft3 or kg/m3) – o, w, g – for oil, water and gas;
 viscosity – a measure of a fluid’s resistance to flow –  (cP, or Pas).
 phase saturations – the fraction of the pore space taken up by a particular phase
(dimensionless) – So, Sw, Sg. Remember that So + Sw + Sg = 1. (In most examples in
this course, we assume Sg = 0.)
 Saturation limits:
o Swi or Swc – the initial water saturation, or connate water saturation,
o Sor – the residual oil saturation.

A reminder of other important reservoir engineering properties is given below in


Table 1.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

Symbol Name Units Meaning/Formulae


Bo, Bw, Formation rb/stb Vol.oil+dissolvedgasinreservoir
Bg volume factors rm3/sm3 Bo 
Vol.oilat STC
for oil, water
and gas
STC = Stock Tank Conditions (60oF; 14.7
psi)
Likewise for water (usually const.) and
gas;
Pb = bubble point pressure, below.

Rso, Rsw Volume of gas at scf/stb Vol. dissolvedgasinreservoir


STC which will (or Rso 
Vol.oilat STC
dissolve in a Mscf/stb)
volume of oil at sm3/sm3
STC, when
subjected to
reservoir Pr and
Temp.

co, cw, Isothermal fluid psi-1 1  k  1  Vk 


ck   
cg compressibilities bar-1
k  P 

Vk  P 
of oil, water and Pa-1
gas
where k and Vk are density and volume
of phase k; k = o, w, g

Table 1
Basic reservoir engineering properties. The units are given first in field units, and
then in metric units. (In metric units, s = surface, r = reservoir.)

©HERIOT-WATT UNIVERSITY December 2017 v2


5

2. Material Balance

2.1 Introduction to Material Balance


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

1. To introduce the central idea of MB and apply it to a simple case which we will
later set up as an exercise for simulation;

2. To demonstrate the complementary nature of MB and reservoir simulation


calculations.

Material balance has been used in the industry for the following main purposes:

1. Determining the initial hydrocarbon in place (e.g. STOOIP) by analysing mean


reservoir pressure vs. production data;

2. Calculating water influx i.e. the degree to which a natural aquifer is supporting the
production (and hence slowing down the pressure decline);

3. Predicting mean reservoir pressure in the future, if a good match of the early
pressure decline is achieved and the correct reservoir recovery mechanism has
been identified.

Thus, MB is principally a tool which, if it can be applied successfully, defines the input
for a reservoir simulation model (from 1 and 2 above). Subsequently, the mean field
pressure decline as calculated in 3 above can be compared with the predictions of the
numerical reservoir simulation model.

Before deriving an example of the MB equations, we quote the introduction of Dake's


(1994) chapter on material balance.

©HERIOT-WATT UNIVERSITY December 2017 v2


6

Material Balance Applied to Oilfields


(from Chapter 3; L. P. Dake, The Practice of Reservoir Engineering, Developments in
Petroleum Science 36, Elsevier, 1994.)

Introduction: It seems no longer fashionable to apply the concept of material balance


to oilfields, the belief being that it has now been superseded by the application of the
more modern technique of numerical reservoir simulation modelling. Acceptance of
this idea has been a tragedy and has robbed engineers of their most powerful tool for
investigating reservoirs and understanding their performance rather than imposing
their wills upon them, as is often the case when applying numerical simulation directly
in history matching.

As demonstrated in this chapter, by defining an average pressure decline trend


for a reservoir, which is always possible, irrespective of any lack of pressure equilibrium,
then material balance can be applied using simply the production and pressure histories
together with the fluid PVT properties. No geometrical considerations (geological
models) are involved, hence the material balance can be used to calculate the
hydrocarbons in place and define the drive mechanisms. In this respect, it is the safest
technique in the business since it is the minimum assumption route through reservoir
engineering. Conversely, the mere act of construction of a simulation model, using the
geological maps and petrophysically determined formation properties implies that the
STOIIP is "known". Therefore, history matching by simulation can hardly be regarded
as an investigative technique but one that merely reflects the input assumptions of the
engineer performing the study.

There should be no competition between material balance and simulation,


instead they must be supportive of one another: the former defining the system which
is then used as input to the model. Material balance is excellent at history matching
production performance but has considerable disadvantages when it comes to
prediction, which is the domain of numerical simulation modelling.

Because engineers have drifted away from oilfield material balance in recent
years, the unfamiliarity breeds a lack of confidence in its meaningfulness and, indeed,
how to use it properly. To counter this, the chapter provides a comprehensive
description of various methods of application of the technique and included six fully
worked exercises illustrating the history matching of oilfields. It is perhaps worth
commenting that in none of these fields had the operators attempted to apply material
balance, which denied them vital information concerning the basic understanding of
the physics of reservoir performance.

Notes on Dake's comments:

©HERIOT-WATT UNIVERSITY December 2017 v2


7

1. The authors of this Reservoir Flow Simulation course would very much like to echo
Dake's sentiments. Performing large scale reservoir simulation studies does not
replace doing good conventional reservoir engineering analysis - especially MB
calculations. MB should always be carried out since, if you have enough data to
build a reservoir simulation model, you certainly have enough to perform a MB
calculation.

2. Note Dake's comments on the complementary nature of MB in defining the input


for reservoir simulation, as we discussed above.

3. Take careful note of Dake's comment on where a reservoir simulation model is


used for history matching. The very act of setting up the model means that you
actually input the STOIIP, whereas, this should be one of the history matching
parameters. The reservoir engineer can get around this to some extent by building
a number of alternative models of the reservoir.

2.2 Derivation of Simplified Material Balance Equations


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

Table 2 below.

The simple example which we will take in order to demonstrate the main idea of
material balance is shown in Figure 1, where the system is simply an under-saturated
oil, with possible water influx. See also Table 3 with the definitions of symbols used.

Note that, although the water and oil are drawn in separate sections in Figure 1, the
material balance equation considers all the water in the “tank” which may be spread
throughout the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

Fluid or formation Compressibility (10-6 psi-1)


Formation rock, crock 3 - 10
Water, cw 2-4
Undersaturated oil, co 5 - 100
Gas at 1000 psi, cg 900 - 1300
Gas at 5000 psi, cg 50 - 200

Table 2
Typical rock and fluid compressibilities (from Craft, Hawkins and Terry, 1991).

Figure 1
Simplified system for Material Balance (MB) with an under-saturated oil above the
bubble point.

©HERIOT-WATT UNIVERSITY December 2017 v2


9

Symbol Definition
N STOIIP – stock tank oil initially in place (stb)
Boi Initial formation volume factor of oil (rb/stb)
Np Cumulative volume of oil produced at time t, pressure, p (stb)
Bo Oil formation volume factor at current t and p (rb/stb)
W Initial volume of water in the reservoir (rb)
Wp Cumulative volume of produced water (stb)
Bw Water formation volume factor (rb/stb)
We Water influx into reservoir (rb) (“e” stands for “encroachment”)
cw Isothermal compressibility of water
crock Isothermal compressibility of rock
p Change in reservoir pressure , p – po

Vf Initial void space (rb); V f  N .Boi 1  Swi  ; W  V f .Swi


Swi Initial water saturation (of the whole system)
cf 
Void space isothermal compressibility (psi-1) c f  1 V f  V f p 

Table 3
Definitions of symbols used in the Material Balance Equation. (rb = reservoir barrels,
stb = stock tank barrels.)

In going from initial reservoir conditions shown in Figure 1 at pressure po to pressure


p, volume changes in the oil, water and void space (rock) occur: Vo,Vw, Vf (Vf = -
Vrock).

The pressure drop is p  p  po . Note that this is current pressure minus initial
pressure. Since the reservoir is depleting p is negative.

The volume balance equation is simply:

Vo  Vw  Vrock  Vo  Vw  V f  0 . (1)

In Appendix A1, we show expressions for each change in volume, and how to derive
the material balance equation. (You are not required to reproduce this derivation.)

From the Appendix, the simplified Material Balance Equation is:

©HERIOT-WATT UNIVERSITY December 2017 v2


10

 Np  Boi Boi  S wi  cw  c f 
   1    p (2)
 N  Bo Bo  1  S wi 

This is for a reservoir with oil and water (no gas), with no water production or water
influx from the aquifer. Also, we have assumed that the pressure remains above the
bubble point.

We can also write this as:

N  Boi Boi  Swi  cw  c f 


1  p     p (3)
 N  Bo Bo  1  S wi 

We can then identify 1 – (Np/N) as the fraction of the initial oil still in the reservoir.
Assuming that we know all the quantities on the right side of Equation (3), we can
__ __
calculate 1 – (Np/N) and plot against  p , as shown in Figure 2. (We take  p so
that it increases along the x-axis.)

Figure 2
1   N p N  
__

Plot of remaining oil,   vs p.

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

©HERIOT-WATT UNIVERSITY December 2017 v2


11

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

Assume that after a certain time, the cumulative volume of oil produced is Np (stb),
__
and that the pressure has declined by an amount  p . This point of depletion is
shown in Figure 4.

Figure 4
Reservoir depletion on a plot following Equation (3).

The value of X (which is actually calculated from the right side of Equation (3) in Figure
 
4 above, is 1  N p N . Since we know Np, we can calculate N. There is an exercise
on this in Tutorial 2.1 for this Topic.

2.3 Conditions for the Validity of Material Balance


The basic premise for the material balance assumptions to be correct is that the
reservoir be "tank like" i.e. the whole system is at the same pressure and, as the
pressure falls, then the system equilibrates immediately. For this to be correct, the
pressure communication through the system must at least be very fast in practice
(rather than instantaneous which is strictly impossible). For a pressure disturbance to
travel very quickly through a system, we know that the permeability should be very
high and the fluid compressibility should be low (pressure changes are communicated

©HERIOT-WATT UNIVERSITY December 2017 v2


12

instantaneously through an incompressible fluid). Indeed, we will show later (Topic 4)


that pressure equilibrates faster – or "diffuses" through the system faster – for larger
values of the "hydraulic diffusivity", which is given by k/(c) (Dake, 1994, p.78).

Dake (1994, p.78), also points out two "necessary" conditions to apply material balance
in practice as follows:

i. we must have adequate data collection (production/pressures/PVT); and

ii. we must have the ability to define an average pressure decline trend i.e. the more
"tank like", the better – large k/(c) – as discussed above.

Although material balance is very useful for understanding general reservoir


behaviour, there are lots of things which material balance cannot do, which is why we
need reservoir simulation. Examples where we need reservoir simulation are:
 Predicting water breakthrough in a water flood;
 Deciding where to drill in-fill wells;
 Complex recovery processes, such as those involved in enhanced oil recovery
(EOR). See Topic 9.

3. Single Phase Darcy’s Law


In this section, we review the single phase Darcy’s Law which will prove to be very
useful when we derive the flow equations of reservoir simulation in Topic 4.

3.1 The Basic Darcy Experiment


Darcy in 1856 conducted a series of flow tests through packs of sands which he took as
approximate experimental models of an aquifer (for the ground water supply at Dijon).
A schematic of the essential Darcy experiment is shown in Figure 5 where we imagine
a single phase fluid (e.g. water) being pumped through a homogeneous sand pack or
rock core. (Darcy used a gravitational head of water as his driving force whereas, in
modern core laboratories, we would normally use a pump.)

©HERIOT-WATT UNIVERSITY December 2017 v2


13

Figure 5
Single-Phase Darcy’s Law. Note that here P = Pin - Pout.

Darcy’s law as given in Figure 5, is in its "experimental" form where a conversion factor,
, is used to allow us to work in various units as may be convenient to the problem at
hand. Note that A is the area perpendicular to flow (i.e. if L is along the x-axis, A is in
the y-z plane). In differential form, a more useful way to express Darcy’s Law and
introducing the Darcy velocity, u, is as follows:

Q k P k  P 
u   .   . 
 A  L   x  (4)

where the minus sign in Equation (4) indicates that the direction of fluid flow is down
the pressure gradient from high pressure to low pressure i.e. in the opposite direction
to the positive pressure gradient.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Consistent
Symbol Dimensions Meaning Units
Darcy field SI-field Proper SI
units

Q L3/T Volumetric cm3/s bbl/day m3/day m3/s


flow rate

L L Length of cm ft m m
system

A L2 Cross- cm2 ft2 m2 m2


sectional
area

 MLT-1 Viscosity cP cP cP Pas

P M.L.T-2 Pressure atm psi bar Pa


(Force/Area) drop

k L2 Permeability Darcy mD mD m2

 Dimension- Conversion 1.00 1.127 8.527 1.00


less factor x10-3 x 10-5

Note: Permeability has dimensions L2 ; e.g. units m2, Darcies (D), milliDarcies (mD);
1 Darcy = 9.869 x 10-9 cm2 = 0.9869 x 10-12 m2  1 m2 .

Table 4
Units used in Darcy’s Law.

Note on Units Conversion for Darcy's Law: the various units that are commonly used
for Darcy's Law are listed in Table 4Error! Reference source not found. above. The
conversion between various systems of units causes confusion for some students.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

Here, we briefly explain how to do this using the examples in the previous figure; that
is, we go from c.g.s. (centimetre - gram - second) units where  = 1 to field units.
Indeed, the Darcy was defined such that  = 1. Starting from the Darcy Law in c.g.s.
units:

k ( Darcy) . A (cm2 )  P (atm) 


Q (cm / s)  1.00
3
. 
 (cp)  L (cm) 

Suppose we now wish to convert to field units as follows:

k (m D) . A ( ft 2 )  P ( psi) 
Q (bbl / day )  ?? . 
 (cp)  L ( ft ) 

How do we find the correct conversion factor for these new units? Essentially, we
convert it unit by unit starting from the c.g.s. expression where we know that  = 1.
We do need to know a few conversion factors as follows: 1 ft = 30.48 cm (exact), 14.7
psi = 1 atm, 1 bbl = 5.615 ft3 = 5.615 x 30.483 cm3 = 1.58999 x 105 cm3, 1 day = 24x3600
s = 8.64 x 104 s. Thus, we now convert everything in the field units to c.g.s. units as
follows (except for cP which is the same):
 P
 psi  
k
 bbl 1.58999x105  (mD) . A ( ft.2 ).30.482 
1000 .  14.7
Q  .  
 day 8.64x10
4
  (cp)  L ( ft.).30.48 
 
Thus, collecting the numerical factors together we obtain:

 bbl   8.64x104 . 30.482  k (mD) . A ( ft.2 )  P  psi  


Q      . 
5
 day   1000 . 1.58999x10 . 14.7 . 30.48   (cp)  L ( ft.) 

which simplifies to

 bbl  -3 k ( Darcy ) . A ( ft. )  P  psi  


2
Q    1.126722x10 . 
 day   (cp)  L ( ft.) 

and hence  = 1.127x10-3 for these units.

©HERIOT-WATT UNIVERSITY December 2017 v2


16

Note that in Table 4, the units in the second last column are actually those used in
the Eclipse simulator. They are not proper SI units. MRST has all variables converted
to proper SI units:
 length in m
 time in s
 pressure in Pa
 viscosity in Pa.s
 permeability in m2.
There are special MRST routines to convert units.

In Appendices A2 and A3, we present an extension of Darcy’s Law into 3D, using vector
and tensor notation. This is included to familiarise the student with the concepts, but
it is not an examinable part of the course. Note, though, that it is useful to be familiar
with the fact that permeability is a tensor quantity, because many textbooks and
papers mention permeability tensors. See, for example, Lie (2016).

In this section, it is suffice to note that, although we usually work in 1D or 2D in this


course, fluid can of course flow in any direction, depending on the direction of the
pressure gradient. Therefore the equations in a simulator must be set up to take
account of this. In a Cartesian model, the flow is given by Darcy’s Law along each axis:

1  P 
ux  - kx  
  x 

1  P 
uy  - ky   (5)
  y 

1  P 
uz  - kz  
  z 

(This is assuming that we have a diagonal permeability tensor – i.e. no sub-grid cross-
flow. See Appendix 3 for full details.)

3.2 Darcy’s Law with Gravity


In the presence of gravity the 1D Darcy Law becomes:

©HERIOT-WATT UNIVERSITY December 2017 v2


17

1  P z 
ux  - kx  - g  . (6)
  x x 

In the case of a simple inclined system with a slope of  degrees, as shown in Figure
6:

 z 
   sin  , (7)
 x 
and
1  P 
ux  - k xx  - g..sin   . (8)
  x 

Figure 6
Model in which the x-direction is inclined to the horizontal.

3.3 Radial Darcy’s Law


In the above discussion, we considered Darcy’s Law in normal Cartesian coordinates (x,
y and z). In Topic 3, we will explain how wells are treated in reservoir simulation.
Because a radial (r/z) geometry is appropriate for the near-well region, it is useful to
consider Darcy’s Law in radial coordinates. In 1D, this simply involves the radial
coordinate, r. In fact, the radial form of the Darcy law can be derived from the linear
form as shown in Figure 7.

©HERIOT-WATT UNIVERSITY December 2017 v2


18

Figure 7
Radial form of the single-phase Darcy Equation.

Notation:
Q = volumetric flow rate of fluid into well
r = radial distance from well
h = height of formation
dP = incremental pressure drop from r  (r + dr) i.e. over dr
A = area of surface at r = 2.r.h
 = fluid viscosity
k = formation permeability

Starting from the radial form of the Darcy Law, as follows:

2 khr  dP 
Q
  dr 
(8)

We can rearrange this to obtain:

Q  dr 
dP     (9)
2 kh  r 

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

Q r  dr  Q r  dr 
r rw

 dP   2 kh r  r  or
r
r dP   2 kh r  r  (10)
w w w

©HERIOT-WATT UNIVERSITY December 2017 v2


19

which gives:

Q  r 
P(r )  ln  
2 kh  rw 
(11)

where we have denoted the radial pressure drop away from an injector (or increase
for a producer) from rw to r as, P(r). Note that, unlike the linear version of Darcy’s
Law, the pressure profile is logarithmic in the radial case. This means that pressure
drops are much higher closer to the well. This is exactly what we expect physically
since the area is decreasing with r as we approach the well and Q is the same;
therefore, the pressure drop, dP, over a given dr is higher. This is shown schematically
for an injector and a producer in Figure 8. The formulae and the ideas developed here
will be used later in Topic 3 on well modelling in reservoir simulation.

Figure 8
Pressure profiles P(r), in radial single-phase flow. Pwf is the well flowing pressure (at
rw).

4. Darcy’s Law for Two-Phase Flow


Darcy's Law was originally applied to single phase flow only. However, in reservoir
engineering, it has been convenient to extend it to describe the flows of multiple
phases such as oil, water and gas. To do this, Darcy’s Law has been modified
empirically to include a term - the relative permeability - which is intended to describe
the impairment of the flow of one phase due to the presence of another. A schematic
representation of a steady-state two phase Darcy type (relative permeability)
experiment is shown in Figure 9, where all of the quantities are defined.

©HERIOT-WATT UNIVERSITY December 2017 v2


20

As in the single-phase Darcy experiment, the fluid permeability for each phase
(sometimes termed the effective permeability to a phase) is:

k f A Pf
Qf  (12)
f L

where Q is volumetric flow rate (e.g. in cc/hr, or m3/day), kf is phase permeability, f is


the viscosity, A is the cross-sectional area of the system (= dy x dz), Pf is the pressure
drop for that phase, L is the length of the system, and f is for fluid = oil or water.

Figure 9
Darcy experiment for two-phase steady-state flow. Oil and water are flowing through
the system at a steady state (i.e. the fluids are flowing at the same rate at the inlet and
outlet). The pressure across the system is measured for each phase.

It is customary to write the phase permeability as the product of the absolute


permeability (permeability when just one phase is flowing) and relative permeability.

kabs krf A Pf


Qf  . (13)
f L

Note that the flow rate, the relative permeability, the viscosity and the pressure drop
are different for each phase. (It may seem strange that the pressure drops are
different, but see Topic 7 on capillary pressure.)

Figure 10 shows an example of relative permeabilities for oil and water. It can be seen
that the relative permeability of a phase depends on its saturation (increasing as the
phase saturation increases). Also, the relative permeabilities are less than 1, because

©HERIOT-WATT UNIVERSITY December 2017 v2


21

the fluid phases interfere with each other. Relative permeability is also explained in
Topic 7.

Figure 10
Example of relative permeability curves for oil and water.
The differential form of the two phase Darcy Law in 1D, again including gravity which
is taken to act in the z-direction, is as follows:

k .krw  Pw z 
uw  -  - g w 
  x x 

k .kro  Po z 
uo  -  - g o 
  x x  (14)

As mentioned above, the flow of the two phases (water and oil, in this case) depends
 Pw   Po 
on the pressure gradient in that phase; i.e. on   and   . The phase
 x   x 
pressures, Po and Pw, at a given saturation, Sw (So = 1 - Sw), are generally not equal.
However, they are related through the capillary pressure, as follows:

Pc  Sw   Po  Pw

More strictly, the capillary pressure is the difference between the non-wetting phase
pressure and the wetting-phase pressure: Pc  Sw   Pnon wett .  Pwett . . We can think of
the capillary pressure as a constraint on the phase pressures. That is, if we know the
capillary pressure function - from experiment, say – then, if we have Po at a given
saturation, we can calculate Pw. (As mentioned above, capillary pressure is explained
more fully in Topic 7.)

©HERIOT-WATT UNIVERSITY December 2017 v2


22

5. Fractional Flow Theory

5.1 Introduction to Fractional Flow


In this section we are going to derive a very important theory for estimating the
behaviour of an immiscible flood (e.g. a water flood). It involves fluid flow in 1D, which
at first may not seem very useful. However, it allows us to make estimates for
breakthrough time for a waterflood and the recovery efficiency at breakthrough.
Although engineers can carry out flow simulations in complex reservoir models, it is
still of great value to be able to do quick “Buckley-Leverett” estimates on a
spreadsheet. (You have actually come across Buckley-Leverett Theory before in
G19RB, Topic 7 on Immiscible Displacement.)

Consider the flow of water and oil in one dimension (1D) and assume that these are
incompressible immiscible fluids. Let water displace oil in a thin homogeneous 1D
porous system of porosity  (Figure 11). The system has a cross-sectional area A, which
is sufficiently small that it is possible to consider the liquid saturations and pressures
to be uniform in any cross-section. In deriving the Fractional Flow, and Buckley-
Leverett theories, we will set up continuity equations for oil and water. We shall use
this sort of approach in Topic 4 for deriving the general flow equations. This means
that mass must be conserved in a system, i.e. the mass of a fluid flowing into a block
of rock within a time interval t minus the mass which flows out during that time is the
mass accumulation. The notation we will use is given in Table 5.

In the calculations below, we ignore gravity and capillary pressure (Po = Pw), so that
Darcy’s Law for the two phases is:

kabs krw . A P k k . A P
Qw   and Qo   abs ro . (15)
w x o x

©HERIOT-WATT UNIVERSITY December 2017 v2


23

Figure 11
Blocks in 1D model used to derive the continuity equation for two-phase flow.

DEFINITION NOTATION

Cross-sectional area of the linear system A

Porosity 
Saturation of water Sw

Connate water saturation Swc

Residual oil saturation Sro

Water saturation at the flood shock front Swf

Viscosity of water and oil w ; o


Relative permeability of water and oil krw ; kro

Fractional flow of water and oil fw ; fo

(fw = Qw /QT ; fo = Qo /QT) (fw + fo = 1)

Derivative of fw with respect to Sw; (fw/Sw) gw

©HERIOT-WATT UNIVERSITY December 2017 v2


24

Distance from injector to producer L

Pressure in water and oil Pw ; Po

Total volumetric flow rate; QT = Qw + Qo QT

Volumetric flow rate of water and oil Qw ; Qo

Density or oil and water o ; w


Total fluid velocity (of oil + water) vT = QT /(A)

Cumulative amount of water injected into W


system at time t (volume)

Table 5
Nomenclature for Fractional Flow Theory. Note that vT is the actual velocity
(interstitial velocity), not the Darcy velocity, which is just QT/A.

Figure 11 shows part of a 1D model. We have split the length into blocks of length
numbered i = 1, 2, … n. We shall use this type of model in deriving the flow equations
in Topic 4. We are interested in block i. Since we are assuming incompressible fluids,
the volume of each fluid which flows into the block within a given time is equal to the
volume which flows out (Figure 11).
Figure 10

The fraction of the flowing stream which is water – the water fractional flow, fw, is:

Qw
fw  (16)
QT

and similarly, the oil fractional flow, fo, is:


Qo
fo   1  fw (17)
QT

From Equation (15):

©HERIOT-WATT UNIVERSITY December 2017 v2


25

Qw   kabs krw A w  P x 
fw   .
QT   kabs krw A w  P x    kabs kro A o  P x 

Cancelling kabs, A and P x gives:

1
krw w 1  k  
fw    1  ro w  (18)
krw w  kro o  kro w   krw o 
1  
 krw o 

5.2 Derivation of the Continuity or Mass Balance Equation


Looking again at Figure 11, consider the mass of water flowing into and out of block i
in time t:

Mass that flows Mass the flows Change in mass


INTO Block i over - OUT OF Block i = in Block i over
time t over time t time t

Mass of water IN over time t =  w  Qw  t x (19)

Mass of water OUT over time t = w  Qw  t x x (20)


Change in mass over time t = A    x  w  Sw t t  A    x  w  Sw t  (21)

Since both phases are assumed to be incompressible the water density,  w , and
porosity,  , are constant and we assume that the cross section area, A , is constant.
From Equations (19), (20) and (21) we get:

 
 Qw xx  Qw x w  t  w  A   x  Sw t t  Sw t   (22)

Divide throughout by Axt.


S w t t  Sw t 

1  Qw
 x x
 Qw x  0 (23)
t A x

©HERIOT-WATT UNIVERSITY December 2017 v2


26

Setting x  0, t  0 gives:

Sw 1 Qw
   0 (24)
t A x

Using the fractional flow notation, where Qw  QT  f w :

Sw QT f w
    0. (25)
t A x

5.3 Buckley-Leverett Solution


Assume that the saturation distribution is a function of position and time, Sw(x,t), the
oil and water viscosities, o and w, are constant, and the relative permeabilities for oil
and water are functions of water saturation, Sw, only. Then the fractional flow, fw, is a
function of Sw alone although Sw is a function of both space and time i.e. Sw (x, t).

Hence:

 f w   df w   S w 
   .  (26)
 x   dS w   x 

and the continuity Equation (25) becomes:

QT  f w   S w   S w 
.      (27)
 A  Sw   x   t 

The continuity equation for the oil phase also has the same form.

The full differential of S w can written as follows

S w S
dS w  dx  w dt (28)
x t t x

©HERIOT-WATT UNIVERSITY December 2017 v2


27

Thus, if x = x(t) is chosen to coincide with a surface of fixed water saturation Sw; i.e. we
imagine that the point x(t) is moving with the constant saturation level, then:

dSw  0
and
 dx   Sw   Sw 
     t   x  (29)
 dt  Sw    

is the rate of advance of the saturation Sw.

Substituting Equation (29) in Equation (27) above reveals that that the following terms
are equal:

 dx  QT  df w 
     (30)
 dt  Sw  A  dS w 

This is the Buckley-Leverett Equation. This equation essentially identifies the velocity
of a given saturation value, Sw, with the derivative of the fractional flow function, fw(Sw)
multiplied by the total fluid velocity (of both oil + water), vT  QT / ( A) .

Integrating with respect to time gives:

W  t   df w 
x  Sw , t     (31)
A  dS w 

where W(t) is the total amount of water which has been injected since t = 0. (It is
assumed that water is injected at the left side where x = 0.)

df w
The coefficient can be evaluated at every saturation, Sw, if the function fw (Sw) is
dS w
known and is differentiable. It follows that if the saturation distribution is known at
time t = 0, it can be found at any other time t > 0.

5.4 Welge’s Method


Figure 12 shows a typical form of the fractional flow fw as a function of water
saturation, Sw. (See Equation (18), which gives the expression for fractional flow as a

©HERIOT-WATT UNIVERSITY December 2017 v2


28

 df w 
function of relative permeability and viscosity.) Since the gradient   increases
 dS w 
and then decreases as a function of Sw, the distance traversed, x, as a function of Sw
(after a given time t) has the form shown in Figure 12. The result shown turned
round in
Figure 13 (left) is obviously physically unrealistic in that saturation, Sw, is a double-
valued function of distance. However, this problem of multiple values in saturation is
eliminated by taking a volumetric balance for the injected water. This means that the
saturation stays constant at the initial saturation, Swc, until then there is an abrupt
increase in water saturation, in the form of a shock front (Figure 13, right).

Welge (1952) analysed this problem and demonstrated that the saturation of the shock
front and the fractional flow at that point can be determined by drawing a tangent to
the fractional flow curve from the Sw = Swc, and f(Sw) = 0. This is known as the Welge
tangent (Figure 14). We do not present the proof here. It is covered in many text
books, such as Dake (1978).

Figure 12
Left: A typical fractional flow curve. Right: the derivative of fractional flow with
respect to saturation, which is proportional to distance gone.

©HERIOT-WATT UNIVERSITY December 2017 v2


29

Figure 13
Saturation vs distance. Left: Same curve as Figure 12, but plotted as S(x). The bold
dashed line shows where the curve is chopped off to give the shock front. Right: The
S(x) curve with the shock front drawn. In these figures Swc = 0.2 and Sor = 0.3
(maximum, Sw = 0.7).

Figure 14
Demonstration of the Welge Tangent

Two other useful measures can be obtained from Welge Theory. Firstly, it can be
shown that the time to water breakthrough is given by:

LA
tbt  . (32)
Qw  df w dS w  Swf

Also, the recovery efficiency (i.e. fraction of the oil which has been recovered) at
breakthrough is:

©HERIOT-WATT UNIVERSITY December 2017 v2


30

1
R . (33)
df w dS w

You do not have to memorize these two equations. They are only valid for a simple
homogeneous 1D system. However, they are useful for rough estimates.

6. Closing Remarks
This Topic forms a link between reservoir engineering and reservoir simulation. Even
if you become an expert in reservoir simulation, you should still use conventional
engineering approaches such as Material Balance (Section 2), and also fraction flow
theory, along with the Buckley-Leverett solution and the Welge tangent.

Some additional derivations are included in the Appendix. They are only there for
reference.

7. References
Amyx, J.W., Bass, D.M. and Whiting, R.L., 1960. “Petroleum Reservoir Engineering”
McGraw-Hill, New York.

Buckley, S.E. and Leverett, M.C., 1942. “Mechanism of Fluid Displacement in Sands”
SPE, doi:10.2118/942107-G.

Craft, B.C., Hawkins, M.F. and Terry, R.E., 1991. “Applied Petroleum Reservoir
Engineering”, Prentice Hall, NJ.

Dake, L.P., 1978. “The Fundamentals of Reservoir Engineering”, Developments in


Petroleum Science, 8, Elsevier.

Dake, L.P., 1994. “The Practice of Reservoir Engineering”, Developments in Petroleum


Science, 36, Elsevier.

Lie, K-A, 2016. “An Introduction to Reservoir Simulation Using MATLAB: User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, www.sintef.no/mrst.

Welge, H.J., 1952. “A Simplified Method for Computing Oil Recovery by Gas or Water
Drive”, SPE, doi:10.2118/124-G.

©HERIOT-WATT UNIVERSITY December 2017 v2


31

8. Appendix
A1 Full Derivation of the Simplified Material Balance Equation

From Section 2.2:


__
The pressure drop (new pressure – initial pressure) =  p  p  po . (A1)
The volume balance is:
Vo  Vw  Vrock  Vo  Vw  V f  0 (A2)
Each of these volume changes can be calculated quite as follows:

 Oil volume change, Vo

Initial oil volume in reservoir = N.Boi (rb)

Oil volume at time t, pressure p = (N - Np). Bo (rb)

Increase in oil volume, Vo = (N - Np). Bo - N.Boi (rb) (A3)

 Water volume change, Vw

Initial reservoir water volume = W (rb)

Cumulative water production at time t = Wp (stb)

Reservoir volume of cumulative water production at time t


= Wp.Bw (rb)

Volume of water influx into reservoir = We (rb)

Water volume change due to compressibility


__
= - W  cw   p (rb)
Increase in water volume,
__
Vw = (W - Wp Bw + We - W  Cw   p ) - W (rb)
__
Vw = -Wp Bw +We - W  Cw   p (rb) (A4)

 Change in the void space volume, Vf

©HERIOT-WATT UNIVERSITY December 2017 v2


32

Initial void space volume = Vf (rb)

Increase in void space volume,


__ __
Vf = (Vf + V f .c f . p ) - Vf  V f .c f . p (rb)

__
Change in rock volume, Vrock = - Vf = - V f .C f . p (A5)

Now adding the volume changes as follows:

Vo  Vw  Vrock  ( N  N p )  Bo  N  Boi


__ __ (A6)
 Wp  Bw  We  W  cw   p  V f  Crock   p  0

Rearranging Equation (A6) and noting that W = Vf.Swi and Vf = N.Boi/(1-Swi), we obtain:

N  Boi  N  Bo  N p  Bo  We  Wp  Bw
 S  c  c f  __ (A7)
 N  Boi  wi w  p  0
 1  S wi 

Equation (A7) is the (simplified) material balance expression for the under-saturated
system given in Figure 1 (as long as it remains above its bubble point).

To illustrate the use of material balance in an even simpler example, let us assume
that there is no water influx (We =0) or production (Wp = 0). Therefore, the MB
equation simplifies even further to:

 S  c  c f  __
N  Boi  N  Bo  N p  Bo  N  Boi  wi w  . p  0 (A8)
 1  S wi 

Note that we can divide through Equation (A8) by N (the initial stock tank oil volume,
stb) to obtain:

Np  S  C  Crock  __
Boi  Bo   Bo  Boi  wi w  p  0
N  1  Swi  (A9)

©HERIOT-WATT UNIVERSITY December 2017 v2


33

which rearranges easily to:

 Np  Boi Boi  S wi  cw  c f 
   1    p (A10)
 N  Bo Bo  1  S wi 

where the quantity (Np/N) is the Recovery Factor (RF) and is the oil recovered as a
__
fraction of the STOIIP. It is seen from Equation (A10) that, at t = 0, Bo = Boi and  p  0
__
and therefore (Np/N) = 0, as expected. Remember also that  p is negative, in
__
depletion. (  p  p  po - new pressure minus initial pressure).

A2 Mathematical Note on Gradient and Divergence


In this Appendix, we review the concepts of gradient and divergence operators. This
is useful because many papers and textbooks use vector notation.

Gradient (or grad) is a vector operation as follows:

  
  i  j  k
x y z (A11)

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

P P P
P  i  j  k
x y z (A12)

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


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

©HERIOT-WATT UNIVERSITY December 2017 v2


34

 P 
 i
 x 
 P 
P   j
y 
 
 P 
 k
 z  (A13)

This is shown schematically in Figure 15, where the three components of the vector
 P   P   P 
P , i.e.   i ,   j and   k are shown by the dotted lines.
 x   y   z 

Figure 15
P showing the components of the vector in the x, y and z-directions.

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

   
.
  i
 x
. y
.
j .
k 
z  (A14)

For example, taking the divergence of the Darcy velocity vector, u, gives the
following:

©HERIOT-WATT UNIVERSITY December 2017 v2


35

 ux i 
     
 . u  i
 x
. y
j . z
k .
 u y j 
 
 uz k  (A15)

where we can expand the right side of the above equation by multiplying out the first
(1x3) matrix by the second (3x1) matrix to obtain a "1x1 matrix" which is a scalar as
follows:

 ux i 
     
 . u  i
 x
. y
j . z
.
k  u y j 
 
 uz k  (A16)
 u u y u z 
 xii 
 x
.
y
j j 
z
.
k k

.
. .
where we use the relationships, i i  j j  k k  1 , to obtain:.
 ux u y uz 
 .u  x  y  z 
  (A17)

Likewise, we can take the divergence of the grad P vector, P , to obtain the
quantity,  . P (sometimes denoted div grad P), as follows:
  P  
  x  i 
  
       P  
 . P   i
 x y
. j .
k   j
z    y  
.
  P  
   k 
  z  
  2 P   2 P   2P  
. .
  2  i i   2  j j   2  k k  .
 x   y   z  
(A18)

©HERIOT-WATT UNIVERSITY December 2017 v2


36

. . .
Again using the relationships, i i  j j  k k  1 , we obtain the familiar
expression:

. P   xP    yP    zP    P


2 2 2
 2 2 2
2

  (A19)

 2 is known as the Laplacian operator:

2 2 2
2   .   div . grad 
x 2

y 2

y 2

In the next section of the Appendix, we will discuss taking the dot product of a tensor
and a vector.

N.B. We now omit the explicit inclusion of the unit vectors, i, j and k in the following
equations.

A3 Darcy’s Law in 3D
You are probably aware that the permeability of a rock may be different in different
directions. This is especially true when comparing the permeability in the horizontal
and vertical directions, i.e. k x  k y  k z . Often the permeability is much lower in the
vertical direction ( kv kh or k z k x ), due to the fact that sediments are usually laid
down in horizontal layers. (See Topic 6.) However, there may be permeability
anisotropy in the horizontal direction.

It turns out that permeability is actually more complex than this. In a heterogeneous
medium, features such as cross-bedding may give rise to crossflow – e.g. flow in the z-
direction due to a pressure gradient in the x-direction (Figure 16). To account for this,
we need a full permeability tensor, k , with 9 terms in 3D (or 4 terms in 2D). A
permeability tensor may be represented in matrix as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


37

 k xx k xy k xz 
 
k   k yx k yy k yz 
k k zy k zz 
 zx (A20)

net flow in z-dir


x

net flow in x-dir z

P

Figure 16
Cross-flow in a heterogeneous model.

Suppose we wish to take a dot product of this tensor, k , with the vector P ; that is
.
k P . The dot product of a tensor and a vector is a vector and the operation is
carried out like a matrix multiplication as follows:

  P  
  x  
 
 k xx k xy k xz  
   P  

k . P   k yx k yy k yz     
k   y  
 zx k zy k zz  
P 
   
  z   (A21)

which multiplies out as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


38

  P    k  P  + k  P  + k  P   
  x     xx  x  xy  
 y 
xz   
 z   
   
 k xx k xy k xz   
 
   P      P 
  P   P   
k . P   k yx k yy k yz       k yx   + k yy   + k yz   
k   y      x   y   z   
 zx k zy k zz    
 P      P         
    k P P
+ k zy   + k zz   
  zx 
       x  
z  y   z  

giving the final result:

   P   P   P   
 k xx   + k xy   + k xz    
   x   y   z   
 
  P   P   P   
k . 
P  k yx   + k yy   + k yz   
   x   y   z   
 
   P   P   P  
 k zx   + k zy   + k zz   
   x   y   z  
(A22)

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

  P  
  x  
 
 k xx k xy k xz  
1    P  
u -
1

k . P  -  k yx

k yy k yz     
  y  
 k zx k zy k zz  
P 
   
  z  

which we may write as:

©HERIOT-WATT UNIVERSITY December 2017 v2


39

   P   P   P   
 ux  k xx   + k xy   + k xz    
     x   y   z   
   
1    P   P   P  
u   u y   - k yx   + k yy   + k yz   
      x   y   z   
   
u     P   P   P  
 z  k zx   + k zy   + k zz   
   x   y   z  
(A23)

and we can identify the three components of the velocity as follows:

1   P   P   P  
ux  - k xx   + k xy   + k xz   
   x   y   z  

1   P   P   P  
uy  - k yx   + k yy   + k yz   
   x   y   z  

1   P   P   P  
uz  - k zx   + k zy   + k zz   
   x   y   z  
(A24)

If the permeability tensor is diagonal i.e. the cross-terms are zero as follows:

 k xx 0 0 
 
k  0 k yy 0 
0 k zz 
 0

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

©HERIOT-WATT UNIVERSITY December 2017 v2


40

1  P 
ux  - k xx  
  x 

1  P 
uy  - k yy  
  y 

1  P 
uz  - k zz  
  z  (A25)

Usually in reservoir simulation, we assume that the permeability tensor is diagonal,


and refer to kxx, kyy and kzz as kx, ky, kz. This makes the equations easier and quicker to
solve. However, sometimes crossflow may be significant, so a full tensor is required.

Also, mathematicians often refer to permeability tensors. For example the MRST book
(Lie, 2016) refers for permeability tensor. MRST is capable of handling full permeability
tensors.

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 3 – Gridding in Reservoir


Simulation
Contents
1. Introduction ............................................................................................................ 3
2. Gridding in Reservoir Simulation ............................................................................ 3
2.1 Introduction .......................................................................................................... 3
2.2 Accuracy of Simulations and Numerical Dispersion ............................................. 6
2.3 Grid Orientation Effects........................................................................................ 9
2.4 Local Grid Refinement ........................................................................................ 10
2.5 Corner Point Geometry ...................................................................................... 11
2.6 Choosing a Reservoir Simulation grid ................................................................. 12
3. Calculation of Block to Block Flows in Reservoir Simulators ................................ 14
3.1 Single-Phase Flow ............................................................................................... 14
3.2 Two-Phase Flow .................................................................................................. 18
4. Wells in Reservoir Simulation ............................................................................... 20
4.1 Introduction ........................................................................................................ 20
4.1 Well Models for Single-Phase Flow .................................................................... 24
4.3 Well Modelling in a Multi-Layer System............................................................. 29
4.4 Modelling Horizontal Wells ................................................................................ 30
5. References ............................................................................................................ 31

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 List and draw different types of grid used in Res Sim.

 Describe what kinds of problem each grid may be used for.

 Explain how numerical dispersion arises.


o Outline how numerical dispersion may be reduced.

 Explain the grid orientation effect.


o Outline how the grid orientation effect may be reduced

 Demonstrate how to calculate the average permeability between two grid cells.
o Also explain which relative permeability should be used when there are
two phases

 Derive an expression for the well index.

 Describe how Peaceman estimated the effective radius in a grid cell.

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
Before learning about the flow equations and how to discretize them for reservoir
simulation, this Topic deals with the related issues of grid selection and well modelling.

Gridding: Examples of various types of grid geometries are presented including 1D, 2D
areal, 2D cross-sectional, 2D radial, and 3D Cartesian cases. A simple description of
two problems associated with gridding, namely numerical dispersion and grid
orientation, is given. Also, some more complex grid structures are discussed.
Treatment of the block to block flows in reservoir simulation is presented and it is
shown how the various inter-block quantities such as the permeability and relative
permeabilities are averaged.
Well Modelling: All interactions between the surface facilities and the reservoir takes
place through the injector and producer wells. It is therefore very important to model
wells accurately in the reservoir simulation model. The central issue with a “well
model” in a simulator is that it must represent a near singular line source within
(usually) a very large grid block.

2. Gridding in Reservoir Simulation

2.1 Introduction
Basically, the gridding process is simply one of chopping the reservoir into a (large)
number of smaller spatial blocks which then comprise the units on which the numerical
block to block flow calculations are performed. More formally, this process of dividing
up the reservoir into such blocks is known as spatial discretisation. Recall that we also
divide up time into discrete steps (denoted t) and this related process is known as
temporal discretisation. The numerical details of how the discretisation process is
carried out using finite difference approximations of the governing flow equations is
presented in Topic 5. However, note that the grid used for a given application is a user
choice - it is certainly not a reservoir “given” - although, as we will see, there are some
practical guidelines that help us to make sensible choices in grid definition.

The simplest type of grid is a Cartesian grid, where all of the grid cells are cuboids, i.e.
each face is a rectangle. The dimensions of the grid cells are denoted by x, y and
z. This grid can be in one dimension (1D), two dimensions (2D) or three dimensions
(3D). Some typical examples of 1D, 2D and 3D Cartesian grids are shown in

Figure 1. The grid cells do not have to be all the same size in a model. For example in
a 3D grid, the thickness of the cells (z) may be varied to reflect the thickness of
geological layers.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

Note that the term “grid” refers to the whole structure. Each unit should be referred
to as a “grid cell” or block. Also note that even in a 1D or 2D grid, each cell is always 3-
dimensional: x, y and z are all non-zero.

 1D linear grids may be used for slim-tube simulations, with a long horizontal
model. Slim-tube experiments are performed in laboratories to test the miscibility
of gas injection into oil. (For example, injection of CO2 into an oil is multi-contact
miscible process, and the CO2 mixes with different components of the oil at
different stages of the flood.) You could also investigate the effect of gravity with
a vertical 1D grid – e.g. gravity stable gas injection.

 2D Cartesian grids may be cross-sectional or areal.


o A 2D cross-sectional (x-z) grid may be used to study a) vertical sweep efficiency
in heterogeneous models; b) the effect of gravity on, for example, a gas flood.
o 2D areal grids may be used to study a) areal sweep efficiencies; b) the stability
of near-miscible gas injection in a thin reservoir layer; c) the benefit of in-fill
drilling in an area pattern flood.

 3D Cartesian grids may be used to study a wide range of production processes.


These grids are very simple to set up. However, nowadays, engineers prefer to use
more complex grids which follow the reservoir architecture.

 Radial Grids
When studying near-wellbore flow, it is useful to construct a radial grid. These may be
1D (r-direction only), 2D (r-z grid), or 3D (r-z- grid, where  is the azimuth angle). From
Topic 2, you know that pressure depends on ln(r/rw) for radial flow, where rw is the
wellbore radius. Therefore engineers usually use a logarithmically spaced grid for the
block size, r. A logarithmic spacing divides up the grid such that (ri/ ri-1) is constant
where ri = (ri- ri-1) as shown in Figure 2. For example, if we take rw = 0.5 ft and the
first grid block size is, r1 = 1ft, then this sets (ri/ ri-1) = 3 since r0 = rw = 0.5ft and r1 =
1.5ft.; hence r2 = 3x1.5ft. = 4.5ft and r2 = 3ft, r3 = 13.5ft and r3 = 9ft, and so on.

©HERIOT-WATT UNIVERSITY December 2017 v2


5

a) x

b) c) x

d)

e)
16
x
y

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

©HERIOT-WATT UNIVERSITY December 2017 v2


6

Figure 2
Radial (r-z) grid geometry.

2.2 Accuracy of Simulations and Numerical Dispersion


Numerical dispersion is essentially an error due to the fact that we use a grid block
approximation for solving the flow equations. Here we present a non-mathematical
description of how numerical dispersion arises, and how it may be reduced.

Consider the simulation of a waterflood in a 1D grid with 5 grid cells, as shown in Figure
3. Each of the sub-figures a) to e) represents a time step, t, of the water injection
process. Block i = 1 contains the injector well which is injecting water at a constant
volumetric rate of Qw, and block i = 5 contains the producer. The system is initially at
water saturation, Swc, and this water is immobile i.e. the relative permeability of water
is zero, krw(Swc) = 0. Each block has constant pore volume, Vp = x.A. where  is the
porosity. In Figure 3a), we see that after time, t = t, some quantity of fluid has been
injected into block i = 1; the volume of water injected is Qw.t and this would cause a
water saturation change in grid block i = 1, Sw1 = (Qw. t)/Vp i.e. the new water
saturation in this block is now, Swc+ (Qw. t)/Vp. Over the first time step, no fluid flowed
from block to block since the relative permeability of all blocks was zero (krw(Swc) = 0).
However, the relative permeability in block i = 1 is now krw(Sw1) > 0. The second time
period of water injection is shown in Figure 3b). Another increment of water, Qw. t,
is injected into block i = 1 causing a further increase in water saturation Sw1. However,
over the second time period, krw(Sw1) > 0 and therefore water can flow from block i =
1 to i = 2, increasing the water saturation such that Sw2 > Swc making the relative
permeability in this block, krw(Sw2) > 0. In the third time step, shown in Figure 3c), the

©HERIOT-WATT UNIVERSITY December 2017 v2


7

same sequence occurs except that fluid can flow from block 1  2 and also 2  3,
because krw(Sw3) > 0. In the fourth and fifth time steps (Figure 3d) and Figure 3e)),
flow can now go from block 3  4 and from 4 5 where, since krw(Sw5) > 0, it can be
produced from this block, although the relative permeability in block 5 will be very
small. Hence, after only five time steps to time t = 5t, the water has reached the
producer in block 5 from whence it can be produced (although not a very high rate
because the relative permeability is very small) and this is an unsatisfactory situation.
If we had taken 10 grid blocks, then clearly a similar argument would apply and water
would be produced after 10 time steps - with an even lower relative permeability in
block i = 10 - and this is more satisfactory. Indeed, this underlies why we take more
grid blocks. This simple illustration explains in a quite physical way the basic idea of
numerical dispersion.

The frontal spreading effect of numerical dispersion can be seen when we try to
simulate the actual saturation profile, Sw(x,t), in a 1D waterflood, and compare this
with the Buckley-Leverett solution presented in Topic 2. There is a shock at the flood
front in the analytical solution, and at time t, the front will have a definite position, x(t).
However, in a grid system with block size x, any saturation front can only be located
within x as shown in Figure 4. If we take more grid blocks (x decreases), then we
will locate the front more accurately. Indeed, taking more and more blocks we will
gradually get closer to the analytical (correct) solution. Hence, one method of reducing
numerical dispersion is to increase the number of grid blocks. An alternative is to use
a numerical method which has inherently less dispersion in it but we will not pursue
this here. Another approach is to use pseudo functions to control numerical
dispersion. These are discussed in Topic 6.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

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

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

©HERIOT-WATT UNIVERSITY December 2017 v2


9

2.3 Grid Orientation Effects


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

Figure 5
Flow between an injector (I) and 2 producers (P1and P2) where the injector-producer
separations are identical but flow is either oriented with the grid or diagonally across
it, illustrating the grid orientation effect.

The effects on the breakthrough time and in the recoveries of these two flow
orientations are shown in Figure 6. The I-P1 orientation tends to lead to earlier
breakthrough and a less optimistic recovery than the I-P2 orientation.

Figure 6
Oil recoveries for aligned and diagonal flows.

©HERIOT-WATT UNIVERSITY December 2017 v2


10

The grid orientation error can be emphasised for certain types of displacement. For
example, in gas injection, the gas viscosity is much less than the oil viscosity (g << o)
leading to viscous fingering instability. This is exaggerated when flow is along the grid
as in the I-P1 orientation.

2.4 Local Grid Refinement


In a reservoir, the changes in pressure, saturations and flows tend to be quite different
in different parts of the system. For example, close to a well which is changing
production rate every day or so, there will be large pressure and saturation changes.
On the contrary, on a flank of the field which is connected to, but is remote from, the
active wells, the pressure may be quite slowly changing and the saturations may hardly
be changing at all. To represent regions with rapidly changing waterfronts will require
a finer grid than will be required for relatively stagnant regions of the system Thus, a
single uniform grid with fixed x, y and z will often not be suitable to represent all
regions of an active reservoir. Instead, the application of some local grid refinement
(LGR) may be much more appropriate.

Figure 7 shows a simple version of a refined grid – not a true LGR, but sometimes
referred to as a “tartan grid”. An example of true LGR is shown in Figure 8, where you
can see that only the area of interest is refined and none of the surrounding area. Note
that in some simulation packages (e.g. Eclipse), you can use radial local grid refinement
to study near-well effects.

Figure 7
A simple version of local grid refinement where the grid is finer in the (central) area
of interest in the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


11

Figure 8
“True” local grid refinement (LGR) where the refined grid is embedded in the coarser
grid.

2.5 Corner Point Geometry


It is common practice nowadays to generate a grid which follows the geometry of a
reservoir. Although there are several different ways of distorting a grid, most
engineers use a Corner Point Geometry grid (CPG) for field studies. In this type of grid
the z-axis is vertical, as in a Cartesian grid, but the cells no longer have a uniform
thickness. CPG grids are complex to build – the coordinates of the corners of each cell
must be specified (hence the name). However, there are packages such as
Schlumberger’s Petrel which can generate CPG grids to be used in simulators. MRST
can read in CPG grids in a similar format to the Eclipse simulator.

Figure 9
Corner Point Geometry

©HERIOT-WATT UNIVERSITY December 2017 v2


12

In Topic 1, CPG grids were shown in Figures 4 and 8, 10 and 22. In Figure 10 below,
we reproduce the example from Figures 4 and 10 of Topic 1, focussing on the grid
structure. Notice how the cells are approximately cuboidal, but are slightly distorted.
Also, this grid has been deformed to represent faulting in the reservoir.

Figure 10
Example of Corner Point Geometry (CPG). This is the MRST SAIGUP example (Lie,
2016).

2.6 Choosing a Reservoir Simulation grid


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

(ii) Grid Geometry/Structure: The next issue is whether we should use a


simple Cartesian grid (x, y, z) or some other grid structure such as a radial
r-z grid, or a corner-point grid (CPG).

(iii) Grid Fineness/Coarseness: How many grid blocks do we need to use?


E.g. is a grid with a few 100 or 1000 grid cells adequate, or do we need a
grid with 100s of 1000s of grid cells?

In Topic 1, we advised that, when considering the complexity of a simulation model,


you should consider the question which has to be answered, or the decision which has

©HERIOT-WATT UNIVERSITY December 2017 v2


13

to be taken. Therefore, the issues of grid dimension, type and fineness should be
appropriate for the question/decision. However, as we will see, there are several
technical considerations that can guide us in these choices.

As mentioned above, a 3D corner-point grid is usually used for a reservoir study. On


the other hand 2D grids may be very useful for sensitivity studies – e.g. a 2D cross-
section to study vertical sweep in a heterogeneous reservoir. Since 2D grids tend to
have fewer cells than 3D grids, a 2D grid will run much faster so many realisations of
the model may be run to investigate uncertainties. Another example of where a 2D
grid might be appropriate is when studying near-well coning, in which case an r-z grid
might be appropriate. 1D grids are not normally used, except for testing or training
purposes. However, as mentioned above, one example where a 1D model is very
useful is for modelling a slim-tube experiment to examine multi-contact miscibility of
CO2 injection into oil.

Often in a field simulation model, the cell thickness (z) is chosen based on the
thickness of geological bedding – e.g. 1-2 m. However, this may give rise to too many
grid cells, so some upscaling (Topic 6) is required. In the horizontal direction, the size
of the cells is usually much larger (~100 m). The main constraint is that there are
several cells between wells, otherwise numerical dispersion will produce early
breakthrough. (If there are still too many cells, a coarse model could be used with LGR
around the wells.)

The grid resolution also depends on the process being simulated. In some cases a
coarse grid may give misleading results. For example, in the case of gas injection, since
gas has a low density compared with oil, the gas will rise to the top of the reservoir
very quickly. This effect may be simulated by assuming vertical equilibrium (VE) in the
reservoir, so that the gas rises instantaneously to the top. On the other hand a fine
grid with lots of layers is required. This is demonstrated in Figure 11. A VE model
allows gas to form a very narrow “tongue” at the top of the model, leading to early
breakthrough and low recovery. (A VE model only has one layer in the z-direction.) If
5 fine-scale layers are used, the tongue will be much fatter, and breakthrough will be
later, leading to better recovery. As the fine-scale model is refined, the tongue will
become thinner and the recovery similar to that of the VE simulation.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Figure 11
Resolving the gas “tongue” in the Vertical Equilibrium (VE) limit in a gas-oil
displacement by increasing the number of grid cells in the vertical direction (from
Darman et al, 1999).

3. Calculation of Block to Block Flows in Reservoir


Simulators

3.1 Single-Phase Flow


In a reservoir simulator, we model fluid flowing from grid cell to grid cell. This happens
in all directions (assuming a 3D grid). However, we will show the calculation of block
to block flow in the x-direction, for two Cartesian grid cells. Also, initially, we will only
consider single-phase flow.

Figure 12
Diagram of two grid cells with different lengths and permeabilities.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

Notation:
Q = volumetric flow rate

xi-1, xi = sizes of blocks (i-1) and i (may not be equal)

x = distance between grid block centres   xi 1  xi  2

ki-1, ki = permeabilities of blocks (i-1) and i (usually not equal)

P = pressure drop between grid block centres.

Applying Darcy’s law to the inter block flow shown in Figure 12 and using the notation
in that figure, we obtain:
 
k A  P  k A  Pi  Pi 1 
Q    , (1)
  x     xi 1  xi 
 
 2 

where k is the average permeability between the two grid cells.

Calculation of Average Permeability


Consider the volumetric flow, Q, and the related pressure drops as shown in Figure 13:

Consider flows from Block (i-1) to the interface of the two blocks, where we denote
the interface pressure as Pf (see Figure 13).

ki 1 A  Pf  Pi 1 
In block (i-1): Q (2)
 xi 1 2

from which we can find the pressure difference (Pf - Pi-1):

Q xi 1  1 
P f  Pi 1    
A 2  ki 1 
. (3)

©HERIOT-WATT UNIVERSITY December 2017 v2


16

Figure 13
Single-phase flow between blocks to determine the correct permeability average to
use for flow simulation.

ki A  Pi  Pf 
In block (i): Q (4)
 xi 2

from which we can find the pressure difference (Pi - Pf):

 
 P  P    QA 2x  k1  .
i f
i
(5)
 i 

Adding Equations (3) and (5) above to eliminate the Pf term gives:

Q  xi 1 xi 
 Pi  Pi 1       (6)
2 A  ki 1 ki 

which rearranges to the following form of the single-phase Darcy Law:

©HERIOT-WATT UNIVERSITY December 2017 v2


17

 
A 2 
Q    Pi  Pi 1  . (7)
  xi 1  xi 
 k 
 i 1 ki 

Darcy’s Law for flow between the two blocks should have the form:

kA  Pi  Pi 1 
Q , (8)
 xi 1  xi
2
where k is the average permeability. (Equation 8 is essentially the same as
Equation 1.)

Comparing identical terms in Equations (7) and (8) above gives:

k 2
 (9)
xi 1  xi xi 1 xi

2 ki 1 ki

which easily rearranges to:

xi 1  xi
k (10)
 xi 1 xi 
  
 ki 1 ki 

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

(i) The appropriate average k is not the arithmetic average, it is the harmonic
average weighted by the grid block sizes.

(ii) This harmonic average gives much more weighting to the lower permeability
value. If the grid sizes are equal, this reduced to the exact harmonic average, kH,
of the permeabilities as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


18

1 1 1 1
    (11)
k H 2  ki 1 ki 

For example, if ki-1 = 100 mD, and ki = 1 mD, kh = 1.98 mD.

Note that because pressure varies from block to block, the viscosity will also vary.
However, viscosity does not usually change much, so we may use the arithmetic
average:

i 1  i
 (12)
2

3.2 Two-Phase Flow


Equation (13) shows Darcy’s equation for two-phase flow, averaged between two
blocks. The absolute permeability is averaged as before, and the viscosity is averaged
for each phase, as above.

kkrp A  Pi 1  Pi 
Qp   . (13)
p x

We need to be careful how the relative permeability is calculated. Consider Figure 14:

In Block i, Swi = (1 – Sor). Only water can flow: krw > 0 and kro = 0.

In Block i + 1, Sw = Swc. Only oil can flow: krw = 0 and kro > 0.

Figure 14
Example with block i at Sor and block i+1 at Swc.

©HERIOT-WATT UNIVERSITY December 2017 v2


19

If we use the arithmetic average to average the relative permeabilities of oil and
water between the two blocks. Then krw  0 which is correct, but also kro  0 ,
which is incorrect.

On the other hand, if we use the harmonic average, both krw and kro will equal zero,
and that is incorrect. (Note you cannot calculate the harmonic average if one of the
permeabilities is equal to zero. However, you will notice that, as one permeability
approaches zero, the harmonic average gets smaller and smaller and also approaches
zero.)

We need to find another way of “averaging” the relative permeability. The solution is
to take the upstream relative permeability. (The upstream cell is the one which is at
higher pressure, so that fluid is flowing from it to the next cell.) We take krw  krw,i .
This is greater than zero, as required. On the other hand, kro  kro,i is zero, which
means that no oil can flow from Block i to Block i + 1.

If the fluid is flowing in the opposite direction (from right to left), we take krw  krw,i 1
which is equal to zero, so no water flows, and kro  kro ,i 1 which is greater than zero,
so oil can flow.

In reservoir simulation, sometimes the surface flows are used, and they have to be
multiplied by the formation volume factor, B, to convert to reservoir flows (i.e. divide
by B on the right side of the equation).

kkr A  Pi 1  Pi 
Q . (14)
B x

As with viscosity, B does not vary much from cell to cell, so the arithmetic average
may be used.

To recap, the following averages are used:


 harmonic average for absolute permeability
 arithmetic average for viscosity and formation volume factor
 the upstream value for relative permeability.

©HERIOT-WATT UNIVERSITY December 2017 v2


20

4. Wells in Reservoir Simulation

4.1 Introduction
The only way fluids can be produced from or injected into a reservoir is through the
wells, so we must include them in our reservoir simulation model. As you may know,
the area of Well Technology is vast and in addition to the long wellbore between the
reservoir and the surface, there are many other technical features of wells that can
have a major impact on the flows into and out of the reservoir. For example, there will
be safety valves at the surface and many different types of completion in the well
construction itself. Here, we will simplify things as much as possible in order to extract
the central functions of the well that we will have to model in the simulator. A
schematic of the total well is shown in Figure 15 where the details of the near well
formation are shown in the inset. The near wellbore flows are thought to be radial in
an ideal vertical well and this will have some relevance in modelling the near-well
pressure behaviour, as discussed in Topic 2 and elaborated upon below. In addition to
these near-well pressure drops, there are several other identifiable pressure drops
between the fluids in the reservoir and the surface oil storage facilities and we may
have to model at least some of these. Indeed, it is this topside pressure behaviour that
“links” or “couples” the surface with the pressure and flows that we are trying to model
in the reservoir using reservoir simulation. The main decision is to determine how
much of the formation to surface well assembly we will actually have to model. The
main pressure drops are shown in Figure 16 (based on Figure 15 of whole well +
pressure drops) and are associated with:

(i) Formation → wellbore flow, Pf w : where fluids flow from a “drainage
radius”, re, at pressure, Pe, to the wellbore. Figure 16 shows the near-well
pressure profile, in the near-sandface region with bottom hole flowing well
pressure (BHFP), Pwf. Thus the formation → wellbore pressure drop, Pf w
is:

Pf w   Pe  Pwf  . (15)

©HERIOT-WATT UNIVERSITY December 2017 v2


21

(ii) The pressure drop, Pwell , that may occur along the completed region of the
wellbore from the bottom of the well (or the “toe” of a horizontal well) to
the wellbore just at the top of the completed interval. In very long wells, this
pressure drop along the wellbore due to friction may be quite significant
although there will be cases where it can be ignored.

Figure 15
Schematic diagram of the fluid flows into a well in a grid block model of the reservoir
through to their storage or export from the field.

(iii) Reservoir → surface pressure drop, Pr s : the pressure drop from the well
at the top of the completed formation just above the reservoir to the
wellhead which is at pressure, Pwh . This P is quite significant and, locally
in any sector of the well, there will be a local pressure drop vs. flow rate/fluid
composition relationship. This may be calculated from models (often
correlations) of multi-phase flow in pipes. As the fluids move up the
wellbore, the pressure drops in oil/water production and free gas may also
appear; thus, we can have three phase flow in the well tubular to the surface
and we may have to incorporate this flow rate/pressure drop behaviour in
our modelling.

©HERIOT-WATT UNIVERSITY December 2017 v2


22

(iv) There will frequently be further pressure drops as the fluids flow from the
wellhead through the surface facilities such as the separators, various
chokes, etc. We will not consider this in detail here although it can be an
important consideration in some field cases e.g. if the well is feeding into a
network gathering system which other wells are also feeding into. This could
be a complex surface gathering system network or a multiple-well manifold
of a subsea production system.

Figure 16
Schematic diagram of the fluid flows from the well through to storage or export
showing the associated pressure drops that occur in the system.

In this section, we will focus on the formation to wellbore pressure drops. Thus, our
main task is to either calculate or set the well flowing pressure (Pwf).

©HERIOT-WATT UNIVERSITY December 2017 v2


23

To set the scene in modelling wells in a simulator, we will first consider a very simple
model well producing only oil into the wellbore. How do we decide what the
volumetric flow rate, Qo, of this well is? We start with a simple case where the
wellhead pressure, Pwh, is set to atmospheric pressure. There is then the additional
reservoir to surface pressure drop, Pr s , to consider. Thus, the well flowing pressure,
Pwf, is given by:

Pwf  Patm  Pr s . (16)

Clearly, if this value of Pwf > Pe (the reservoir pressure), then no oil can be produced.
However, if Pwf < Pe, then some oil will flow into the well and we can now calculate
how much. As we will see, this will depend on the physical properties of the system
such as the permeability of the rock, the viscosity of the oil, the precise geometry of
the well etc. However, in our simple conceptual well, we will take all of these
quantities as “givens” for the moment. Suppose the well does flow at a volumetric
flow rate, Qo, for reservoir pressure, Pe, and well flowing pressure, Pwf. We can then
define a productivity index, PI, of the well as follows:

Qo  PI  Pe  Pwf  , (17)

where possible units of PI could be bbl/day/psi, for example. The above equation
basically states the oil production rate per psi of drawdown. This simple equation takes
us back to our original question on “what/who decides on Qo?”. In our simple case,
the answer is now clear; i.e. some things are “givens” - e.g. PI and Pe in a virgin oil
producing system - and some we can set within limits - e.g. Pwf by setting the wellhead
pressure. However, you may be able to set a flowrate, Qo, by installing a downhole
pump (an ESP - electrical submersible pump). In that case, you would set Qo and then
calculate Pwf where we are still considering the reservoir pressure (Pe) as a given. But
clearly we cannot set Qo to any arbitrarily large value since the lowest possible value
of Pwf = 0 (a vacuum), which would then set a maximum value of Qo given by:

Qo,max  PI  Pe (18)

So, in summary, we can set either a pressure or a flowrate but (a) not both and (b) in
either case, within limits.

©HERIOT-WATT UNIVERSITY December 2017 v2


24

But, can’t we affect the well PI or the reservoir pressure, Pe? We can actually change
the PI of a well by “stimulating” it possibly by locally hydraulically fracturing the well
or by acidising it to increase the effective permeability of the near well region. In
addition, we can increase (or more commonly maintain) the reservoir pressure (which
relates to the “reservoir energy”) to some extent by injecting a fluid - usually water or
gas in another injector well. However, the basic well controls are either setting
pressure or flow rate and this must be kept in mind when we model wells in reservoir
simulation. We elaborate on these ideas in the following section where we introduce
the central idea of a well model.

4.1 Well Models for Single-Phase Flow


We now consider how a well model can be developed, firstly in our simple conceptual
reservoir producing only oil. Figure 17 shows the local pressure profiles in a simple
homogeneous well system in single phase flow (see Topic 2, Section 3.3). The pressure
profile close to the wellbore, assuming radial flow, was derived in Topic 2 and is given
by:

Q r 
P  r   ln   , (19)
2 kh  rw 
where h is the thickness of the formation, and rw is the well bore radius.

Taking the pressure at radius re as being the reservoir pressure, Pe, then gives:

Q r 
P  re    Pe  Pwf   ln  e  . (20)
2 kh  rw 

©HERIOT-WATT UNIVERSITY December 2017 v2


25

Figure 17
Pressure distribution near a well: a) near-well pressure profile for a radial system; b)
corresponding quantities in a Cartesian grid block.

Equation (20) can be re-arranged to give:

2 kh
Q P  P  .
 ln  re rw  e wf
(21)

Hence, from Equation (17) above, we can identify the productivity index of the well
as:

2 kh
PI  . (22)
 ln  re rw 

This now demonstrates exactly how the quantities k, h, , rw and re affect the well
productivity. All of these factors behave as we might expect them to physically e.g. as
k, PI; as , PI etc.

Now consider how this relates to the pressures in the simulation block shown in Figure
18 and Figure 19. In a grid block, the pressure is thought of as being constant
throughout the block although we know that it should be varying continuously across
the block; we will refer to this as the average block pressure. The size of the grid block
in our example is (x, y) and, for simplicity, we will assume that, x = y. Looking at
the expression for PI in Equation (22) and the quantities we have in the grid block, it is
easy to make direct relations for some of them - obviously k,  and h and also possibly
rw and Pwf, although these latter two do not seem to appear in the grid model. The
drainage radius, re, and the reservoir pressure, Pe, which appear in the radial model do
not appear in the grid model - instead, the block size (x, y) and average block
pressure ( P ) appear. This immediately suggests the following 2 questions:

1. what is the relation between re , and the block size (x, y)?

©HERIOT-WATT UNIVERSITY December 2017 v2


26

2. what is the relation between Pe and the average grid block pressure, P ?

Figure 18
Schematic diagram to show how the near-well pressure relates to the corresponding
quantities in a Cartesian grid block.

Figure 19
The relationship between re and the block dimensions, x and y.

The issue is defined quite clearly in Figure 18. From this figure, we would like to choose
re such that Pe coincides with the average grid block pressure, P . The latter quantity
( P ) is calculated in the simulation itself. In fact, we need to know how to calculate re
from the quantities x and y as indicated in Figure 19 where we show the re as
function of x and y, i.e. re(x, y).

If we know the formula for re(x, y), then we can calculate the PI (Equation 22) and
use this in the simulation to couple the quantities Qo (oil flow rate) and P (average
grid block pressure); i.e.

©HERIOT-WATT UNIVERSITY December 2017 v2


27


Qo  PI P  Pwf  (23)

Here, we can set either Qo or Pwf and then calculate the other one from P (and the
known PI). This was achieved in a very simple but ingenious way in a classic paper by
Donald Peaceman (1978), a pioneer of numerical reservoir simulation. Peaceman did
this by carrying out a 2D numerical solution of the pressure equation on a Cartesian (x,
y) grid for a quarter five-spot configuration as shown in Figure 20. From Equation (20),
we know that:


 r  2 kh P  r   Pwf
ln   
 (24)
 rw   Q

Hence, if we plot the pressure at grid blocks away from the well block vs. the well block
spacing on a logarithmic scale as shown in Figure 21, then we can extrapolate back to
find the equivalent radius where P = Pe in terms of the well block dimension (x). It
turns out that the simple relation is (for x = y):

re  0.2x . (25)

Therefore, we have a very simple way of calculating the PI of a well in a simulation grid
block. The well PI is also sometimes known as the “well connection factor”,
“connection transmissibility factor”, or the “well index”, Iw.

The simple Peaceman formula applies to a well in a radial environment (the five-spot
configuration is as close as we can get to radial using a common 2D Cartesian grid) and
for x = y. In fact, some modification to the simple formula is required for wells in
“corner” locations or set close to a boundary. Also, if x  y, but the well is isolated
(radial flow), then:

re  0.14 x 2  y 2 . (26)

©HERIOT-WATT UNIVERSITY December 2017 v2


28

Figure 20
The 2D areal grid used to compare pressures with the expected radial profiles
(Peaceman, 1978).

Figure 21
Plot of pressures from the 2D areal grid vs log(r), used to find re (from Peaceman,
1978).

When two phases are present, the equation for flow at a well is:

2 kkrp h
Qp 
 p ln  re rw 
P  P 
e wf (27)

©HERIOT-WATT UNIVERSITY December 2017 v2


29

where the subscript “p” stands for phase (oil, water or gas). Remember that the
relative permeability and the viscosity are different for each phase. In this case, the
well PI, or well connection factor is sometimes given as:

2 kh
PI  , (28)
ln  re rw 

missing out the viscosity, so that the PI does not depend on fluid properties. Then the
well flow is:

PI  Pe  Pwf  .
krp
Qp  (29)
p

4.3 Well Modelling in a Multi-Layer System


Usually a well is connected to more than one grid cell, and a PI, or well connection
factor must be calculated for each grid cell. For example, if we have a model with 4
layers, numbered k = 1 to 4, the well connection factor for one layer (one connection)
is:

2 kk hk
PI k  , (30)
 re 
ln  
 rw 

because the permeability of each layer may be different, and also the thickness. See
Figure 22. (Note that we’ve assumed here for simplicity that the cells are the same
size in the horizontal so that re does not vary.) The total flow of phase, p, will be:

 
4 krp
Qp   PI k P k  Pwf . (31)
k 1 p

©HERIOT-WATT UNIVERSITY December 2017 v2


30

Figure 22
Two-phase flow into a well in a multi-layered system.

4.4 Modelling Horizontal Wells


Figure 23 shows the trajectory of a “horizontal” well in a reservoir simulation model.
This is not well represented by the purely radial r/z model grid discussed above in the
context of a vertical well. Hence, it is less likely that the well connection factors
calculated as shown in previous sections will apply for a horizontal well. However, the
basic principle is very similar. That is, each well sector intersects a grid block (i,j,k) even
although the well may be going through this block in, say, the x (i) direction. The flows
between the well sector and the grid are given by an expression of the form:

 
Qi , j ,k  PIi , j ,k Pi , j ,k  Pwf ,i , j ,k , (32)

where (i, j, k) are the coordinates of the cell connecting to the well. The actual form
of the productivity index expression, PIi,j,k , may be rather more complex since (a) the
well may intersect the block in a more complex way and (b) the aspect ratio of the
block is rather different when a horizontal well intersects it in that the x-direction well
is very close to the z-boundaries since z is often smaller than x or y.

©HERIOT-WATT UNIVERSITY December 2017 v2


31

Figure 23
Section of a 3D reservoir model showing two horizontal wells, plus two vertical wells.

5. References
Darman, N.H., Sorbie, K.S. and Pickup, G.E., 1999. “The Development of Pseudo
Functions for Gravity-Dominated Immiscible Gas Displacements”, SPE 51941,
presented at the SPE Reservoir Simulation Symposium, Houston Texas, 14-17 February,
1999.

Lie, K-A, 2016. “An Introduction to Reservoir Simulation Using MATLAB. User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, www.sintef.no/mrst/.

Peaceman, D.W., 1978. “Interpretation of Well-Block Pressures in Numerical Reservoir


Simulation”, SPEJ, June 1978, 183-194.

©HERIOT-WATT UNIVERSITY December 2017 v2


32

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 4 – Derivation of the Flow


Equations
Contents
1. Introduction ............................................................................................................ 3
2. The Single-Phase Pressure Equation ...................................................................... 5
2.1 The Physics of Single-Phase Compressible Systems ............................................. 5
2.2 The Single-Phase Pressure Equation .................................................................... 7
2.3 The Simplified Compressible Pressure Equation ................................................ 13
2.4 Extension of the Single-Phase Pressure Equation to 2D and 3D ........................ 16
3. The Two-Phase Flow Equations ............................................................................ 19
3.1 Review of Two-Phase Flow Concepts ........................................................... 19
3.2 Derivation of the Two-Phase Conservation Equations ................................. 20
3.3 The Simplified Two-Phase Pressure Equation .................................................... 24
3.4 The Full Two-Phase Pressure Equation .............................................................. 27
4. Closing Remarks.................................................................................................... 28
5. References ............................................................................................................ 28
6. Appendix to Topic 4: The Derivation of the Full Two-Phase Pressure Equation .. 28

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 State the two main steps in deriving the pressure equations.

 Derive the single-phase pressure diffusion equation in 1D.

 Explain what is meant by a non-linear partial differential equation.

 Derive the simplified linear pressure equation.

 Derive a simplified two-phase flow equation in 1D.

 Explain how the two-phase pressure equation is comprised of viscous, capillary


and gravity terms.

Note that there are some lengthy derivations in this Topic. You are expected to know
the derivations in the main part of this Topic, but not the ones in the Appendix.

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
The central activity of reservoir simulation is the numerical solution of the multi-phase
flow equations in real reservoir systems. This chapter introduces the underlying
equations of flow through porous media which are solved in a reservoir simulation
code. We start with the single-phase flow equation for a compressible system. The
mathematics for multi-phase fluid flow is not much more complicated. However, there
are many terms in the equations (especially in a 3D system!), so we shall make some
simplifying assumptions when dealing with two-phase flow.

The resulting equations are, in general, non-linear partial differential equations (PDEs),
which cannot usually be solved analytically. In the next topic, Topic 5, we show how
to discretize these equations for numerical solution in a reservoir simulator.

General Approach
The general approach for deriving the flow equations is basically the same for both
single- and two-phase flow. We first derive a mass balance between the flows and the
accumulation (local mass build-up or decline) in a local control volume. A control
volume can be thought of as a typical isolated grid block in the system (Figure 1). Note
that the mass balance equation must be correct: it is simply a mathematical way of
expressing something that is a fact. Mass balance simply states that over a given period
of time (t, say):

Mass that flows in – Mass that flows out = Mass accumulation.

Usually, we do not think of flow in terms of mass, but in terms of the volumetric flow
rate, usually denoted by Q, in units of m3/day, or bbl/day. So the mass flow into or out
of a block in time, t, is given by:

m = Qt,

where  is the density (e.g. kg/m3 for metric units). We then use Darcy’s Law to relate
Q to the pressure gradient P x . Hence we derive pressure equations.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

Figure 1
Basic Principles of Mass Balance and Darcy’s Law as the basis for all equations for
flow through porous media.

In Section 2, we start by deriving the single-phase flow equation for a compressible


system. This is essentially a pressure equation, since this is the only quantity we need
to find. The pressure distribution in space is the main unknown in the system and we
need to find this as a function of time as the system evolves. Pressure is denoted by
P(x, y, z, t) (for a 3D system). If we have a way of finding P(x, y, z, t), we can then find
the flow rates from the pressure gradients; i.e. we use Darcy’s Law.

Although we can obtain the single-phase pressure equation for a compressible


fluid/rock system in 1D, it turns out to be a non-linear partial differential equation
(PDE). There is no known analytical solution to this equation for the general case. That
is, even the apparently simple single-phase pressure equation cannot be solved using
“well known” mathematical functions (as discussed in Topic 1). There are two things
which we can do: (i) use a numerical method to solve the equations; or (ii) simplify the
equations by making various assumptions that allow us to solve the equation
analytically. We take the latter course of action in this Topic, and then study the
numerical solution of the equations in Topic 5. Section 2 describes how we simplify
the full equation for a compressible fluid in order to obtain an equation which we can
solve. Indeed the simplified equation that arises is the main equation used in single
phase well testing.

In Section 3, we derive a simplified version of the two-phase flow equations. However,


we also show the full two-phase flow equation for a 1D system, for completeness. In
two-phases, we have saturation equations in addition to pressure equations.

©HERIOT-WATT UNIVERSITY December 2017 v2


5

2. The Single-Phase Pressure Equation

2.1 The Physics of Single-Phase Compressible Systems


Before deriving the single-phase pressure equations, consider first the “physics” of
what is happening in a compressible single-phase system. Figure 2 shows a long thin
“reservoir” containing compressible fluid and rock, which we can consider as being a
1D system. The fluid and rock have compressibilities, cf and cr (where cf >> cr). Imagine
that the “reservoir” is horizontal and has two wells in it, at either end as shown in
Figure 2a). Well 1 is located on the left, at x = 0, and Well 2 on the right, at x = L.
Assume that the wells have both been shut in, so that they have reached a steady state
– i.e. the pressure is constant across the system and equals Po (ignoring the
gravitational potential, since the reservoir is thin in the z-direction). This is shown as a
dashed line (at t = 0) in Figure 2b) where the pressure profile across the reservoir is
P(x) = Po.

Now consider what happens to the pressure profile if we cause the pressure to rise
suddenly at Well 1 by injecting fluid (identical to the fluid already in the reservoir).

Is the pressure disturbance immediately “felt” at Well 2 at x = L? The answer is, of


course, “no” since the pressure disturbance will take some time to be transmitted from
Well 1 to Well 2. A short time after this disturbance at Well 1, say at time t1, the
pressure profile will be P(x, t1) as shown in Figure 2b). Similarly, at a later time, t2 , it is
shown as P(x, t2).

But what happens at Well 2? This depends on what we do with Well 2. Suppose we
leave it shut in, what will eventually happen? If we were injecting a volumetric rate,
q1(t) into Well 1, such that it maintained a constant pressure P1(x=0, t), we should be
varying (reducing) q1 with time, the outcome is predictable. We would keep pumping
in at (decreasing) rate q1(t) until we had “blown up” the pressure in the reservoir – like
pumping up a car tyre – to a pressure P1. Hence the value of q1 would reduce to zero,
and the reservoir pressure in both wells would settle to P(x,t) = P1.

©HERIOT-WATT UNIVERSITY December 2017 v2


6

Figure 2
a) Schematic diagram of a long thin “reservoir” containing compressible fluid and
rock.
b) Pressure profile along the system at various times from t = 0, t1, t2, etc. This shows
the pressure diffusing across the system.

On the other hand, suppose we simply pumped fluid into Well 1 at a constant rate q1
(now fixed) and we held the bottom hole pressure of Well 2 to be constant at P(x, L) =
Po (the initial pressure). What would happen in this case? With a little thought, you
can probably visualise the sequence of events, as follows:

Firstly, at early time, up to t2, and a little later, nothing will happen at the Well 2. It
would still have pressure Po and would not be flowing;

The pressure “wave” (actually not a wave, although people tend to refer to it as this)
would propagate or diffuse through the reservoir as shown in Figure 2b); the “speed”
of this wave is governed by the diffusivity which is quantified by k c f  as will be
shown later. Note that the larger the fluid compressibility (cf) or porosity (), the
“slower” the wave will propagate;

At a certain time, denoted by t3, the pressure disturbance just reaches the producer,
Well 2, as shown in Figure 2b);

©HERIOT-WATT UNIVERSITY December 2017 v2


7

At t = t3, Well 2 must start to flow in order to maintain the pressure at P = Po (this works
like a back-pressure regulator);

As time proceeds ( t   ), the pressure wave starts to be felt all across the reservoir
such that the flow rate at Well 1 is (still fixed at) q1, and the pressure settles to a
constant value P1;

Likewise, as t   , Well 2 (still fixed at Po) will go to steady-state flow exactly at


flowrate q1;

At t   , the pressure profile P(x, t   ,) will tend to a straight line, as shown by


the inclined dashed line in Figure 2b).

This “thought experiment” is extremely useful to appreciate the physics of flow in a


compressible system. It also introduces the idea of boundary conditions. That is, not
only do we need a way of modelling the development of pressure in a reservoir (i.e.
finding an equation for P(x, t)), we also need to appreciate what boundary conditions
to apply in a given situation. In this case, we have identified the boundary conditions
as well constraints. Well1 is a rate-constrained injector and Well 2 is a pressure-
constrained producer.

In Tutorial 1 for this topic, you are asked to describe the sequence of events which
would occur in the simple 1D system with varying properties and boundary conditions.

2.2 The Single-Phase Pressure Equation


We now return to the task of deriving the single-phase flow equations for a
compressible system. We refer to the control volume which is shown as block i in
Figure 3. We have divided up the x-axis into increments of (constant) size, x, and
constant cross-sectional area, A (= y.z). Flow is considered to be in the positive x-
direction (i increasing). The fluid has density, , which may depend on pressure, i.e.
(P). The porosity is denoted by  and it too may depend on pressure, (P); in block i,
the porosity is i. The volumetric flows across the boundaries of block i are given by:
qi-1/2 and qi+1/2 as shown in Figure 3. The dimensions of the flows q are volume/time,
and typical units may be stb/day, or m3/s.

Now we apply mass conservation to block i:

(1)

©HERIOT-WATT UNIVERSITY December 2017 v2


8

We can write mathematical expressions for each of the terms in Equation (1) as
follows:

(2a)

Similarly, for the mass flowing out of the block:

(2b)

Figure 3
Control volume for the application of material balance for the single-phase flow
equation.

©HERIOT-WATT UNIVERSITY December 2017 v2


9

The change in mass in block i may therefore be calculated as:

(3)

Note that we have changed signs in the last step for reasons that will become apparent
later on. The density  calculated at location (i-1/2) is an average of the density in
blocks (i-1) and i. (Similarly for the density at (i+1/2).)

Now we use an alternative way of expressing the change in mass of block i.

Therefore:

(4)

Note that the size of the block is assumed constant – i.e. xA is constant. See Figure
4.

©HERIOT-WATT UNIVERSITY December 2017 v2


10

Figure 4
The expression for the mass of a fluid in a grid block.

Now apply the mass balance equation (Equation 1) by equating the expressions in
Equations (3) and (4).

  t t    t  xA    q i 1/2   q i 1/2  t (5)

Divide through Equation (5) by constants xAt:

 q    q  
 A    
  t t    t    i 1/2  A i 1/2 
 , (6)
t x

where we have taken the constant area A into the inner parenthesis of Equation (6).

Note that q A is the Darcy velocity, u. So Equation (6) becomes:

  t t    t   u  i 1/2   u  i 1/2 


 . (7)
t x

©HERIOT-WATT UNIVERSITY December 2017 v2


11

Note that this is an exact equation, because mass balance is exact. We can take the
limits of Equation (7) as t  0 and x  0 to obtain the equivalent differential
equation.

  t t    t  (  
lim   (8)
t 0 t t

and

 u  i 1/2   u  i 1/2   u 


lim    . (9)
x 0 x x

Therefore

     u 
 . (10)
t x

Equation (10) is the differential form of the conservation equation and remember that
it is an exact equation. Note that there is a symmetry between  and . These two
quantities appear in an identical manner on the left side of this equation. Therefore,
if the mass in a grid block stayed the same with time, i.e.     t  0 , then this
could be because the fluid density decreased (fluid expanded) and the rock porosity
increased (the rock contracted or compressed).

We now assume that Darcy’s law holds for u, that is:

k  P 
u . (11)
  x 

We substitute this into Equation (10) (taking care of the signs) to obtain:

       k  P  
    (12)
t x    x  

This equation is now inexact, because Darcy’s Law is just an empirical equation.
However, it is necessary to use a flow law such as Darcy’s Law to provide a link between
fluid velocities and pressure gradients. Equation (12) still does not have pressure (P(x,
t)) explicitly on the left side. However, we know that  and  both depend locally only
on pressure, i.e. they can be written (P) and (P). We can therefore manipulate the
left side of Equation (12) as follows, using the chain rule of differentiation:

©HERIOT-WATT UNIVERSITY December 2017 v2


12

         P 
   (14)
t P  t 

where we consider the term     P as a generalised fluid and rock compressibility


term, C(P). Equation (14) then becomes:

  P     k  P  
C  P     . (15)
t x    x  

Some points to note about Equation (15):

It is a non-linear partial differential equation (PDE). By “non-linear” we mean that the


coefficients in the equation, the “input”, depend on the quantity we are trying to find,
i.e. the unknown pressure P(x, t). In other words quantities such as C(P), (P), (P)
etc depend on the answer. Such equations are difficult to solve analytically for general
cases, but can usually be solved by one of two approaches:

 By solving them numerically where we can handle the non-linearities using certain
types of iterative methods (Topic 5).

 We may simplify the equation to the extent that it becomes soluble.

Both of these approaches gives an approximate solution. In the numerical approach


the full equations are approximated using a method such as the finite difference
method (Topic 5). In the analytical approach, an exact solution may be obtained for
the equations, but during the simplification of the equations, some physics may have
been “lost”. In the next section, we take this second approach.

Going back to Equation (10), we could include gravity effect by taking the Darcy
equation with gravity as follows.

k  P z 
u   g  (16)
  x x 

(instead of Equation (11)). If the 1D system has a constant incline z(x), then
z x  sin   where  is the angle of the incline (Topic 2).

k  P 
u   g sin    (17)
  x 

The generalised form of the single-phase equation is then:

©HERIOT-WATT UNIVERSITY December 2017 v2


13

  P     k  P 
C  P      g sin     . (18)
t x    x 

2.3 The Simplified Compressible Pressure Equation


In this section, we take Equation (15) as our starting point and introduce some
simplifying assumptions. The aim is to obtain a simpler equation which can be solved
analytically for given boundary conditions.

In fact, to make clear what is happening to the left side of the equation, we start with
Equation (15) as follows:

        P     k  P  
      (15)
 P  t x    x  

We first list the assumptions which we will make to the above equation, and then we
go on the examining their consequences.

Simplifying Assumptions
1. Viscosity, , is constant (with x and P);
2. Permeability and porosity, k and , are constant (with x and P), i.e. the system is
homogeneous and the rock is incompressible.
 P 
2

3. The pressure gradients,  P x  , are “small” such that   0.


 x 
1   
4. The fluid has constant compressibility, cf, i.e. c f    constant .
  P 

Assumption 1 is probably quite reasonable since viscosity does not vary greatly for
model oils (or water) over small pressure ranges. Assumption 2 is quite drastic since it
says that the permeability and porosoity (k, ) are constant throughout the reservoir.
However, for a real system permeability especially is often very heterogeneous. The
second part of Assumption 2 is that the rock is incompressible, and this is quite
reasonable, because usually cf >> crock. Assumption 3 is rather odd: it is clearly designed
to get rid of “difficult” terms like  P x  . Assumption 4 – fluid compressibility, cf,
2

does not vary with pressure, is again quite reasonable for most (non-critical) reservoir
oils.

Apply Assumptions 1 and 2 to Equation (15):

©HERIOT-WATT UNIVERSITY December 2017 v2


14

   P  k   P 
     , (19)
 P  t   x  x 

which rearranges to:

    P    P 
     . (20)
k  P  t  x  x 

We now expand the right side of Equation (20) as follows, using the product rule:

  P     P   2P 
        2 . (21)
x  x   x  x   x 

The density, , is only a function of pressure, i.e. (P), so we can expand this equation
using the chain rule:

  P     P  P   2P 
         2 , (22)
x  x   P  x  x   x 

which gives:

  P     P   2P 
2

        2 (23)
x  x   P  x   x 

Assumption 3 can now be used to eliminate the first term on the right side of
 P 
2

Equation (23), i.e.    0 . Therefore


 x 

     P   2 P 
       2 . (24)
 k  P  t   x 

Dividing both side of Equation (24) by  we get:

   1    P    P 
2

      2  . 
 k    P  t   x 

©HERIOT-WATT UNIVERSITY December 2017 v2


15

1
But, note that c f    P  , so we have:

  c f   P    2 P 
     2  (25)
 k   t   x 

This equation is more commonly written as:

 P   k   2 P 
    2  , (26)
 t    c f   x 

k
where the constant is usually referred to as the hydraulic diffusivity, Dh  .
 c f

Equation (26) is now a linear PDE and has the form of a linear diffusion equation:

 P   2 P 
   Dh 2 
. (27)
 t   x 

Note that following about the above equations (26 or 27):

i. They are simplified (slightly) compressible 1D flow equations in a linear


porous medium, i.e. in Cartesian form.

ii. Analytical solutions are available for a range of boundary conditions. (See,
for example, Crank, The Mathematics of Diffusion, 2nd edition, Oxford,
Clarendon Press, 1975).

iii. In its radial form (i.e. in radial coordinates, r, rather than x) Equation (27)
becomes the following:

 P  Dh   P 
  r , (28)
 t  r r  r 

which is the well-known equation of well testing. Likewise, this equation has
a number of well-known analytical solutions for various boundary conditions.

©HERIOT-WATT UNIVERSITY December 2017 v2


16

iv. The reason these equations have many ready-made analytical solutions
available is because these diffusion equations are well-known and are identical
in form to the equations of heat conduction which have been studied for many
years (Carslaw and Jaeger, 1959).

2.4 Extension of the Single-Phase Pressure Equation to 2D and 3D


For the 2D case, the control volume is now grid block (i.j) as shown in Figure 5. The
flows are shown at the boundaries as before and are labelled as (i – ½) for flow from (i
– 1, j) to (i, j) in the x-direction, and (j + ½) for flow from (i, j) to (i, j + 1) in the y-
direction, etc. The areas perpendicular to flow are given by Ax for flow in the x-
direction and Ay for flow in the y-direction. Note that:

Ax  yh Ay  xh

where h is the model thickness (= z) (Figure 5). In addition, we show a source/sink
term due to a (single-phase) well. This well injects or produces exactly the same fluid
as is in the reservoir already. The well flow rate into block (i, j) is described by qwij,
which is the volumetric flow (e.g. in units of m3/day).

Mass flow rate into or out of block (i, j) through a well = qwij ij . (29)

When calculating the change in mass due to the flow into/out of the control volume
we now need to consider flow in the x-direction, the y-direction and the well.

Change in mass in block (i, j) over time t due to flows across the boundaries and the
well:

(30)
The accumulation term is the same as before (Figure 4):

Change in mass in block (i,j) between time t and t+t

 xyh t t  xyh t   t t   t  xyh (31)

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Figure 5
A 2D, x-y grid showing the control volume for calculating mass balance.

We then equate expressions in Equations (30) and (31) to obtain:

 t t   t  xyh

   q  i 1/2   q  i 1/2  t   q   j 1/2   q   j 1/2  t   qw  ij t


 
(32)

Note that the signs have been adjusted on the right side. Now dividing through by
xyht gives:

 t t   t   q  i 1/2   q  i 1/2   q   j 1/2   q   j 1/2 


t

xAx

yAy

 qw  ij

(33)

©HERIOT-WATT UNIVERSITY December 2017 v2


18

where we have used xyh = xAx = yAy. (See Figure 5). Also, we have set
qw  qw  xyh  for convenience. We know that q/A is the Darcy velocity, so
Equation (33) becomes:

 t t   t   u  i 1/2   u  i 1/2   u   j 1/2   u   j 1/2 


t

x

y
 
 qw 
ij

(34)

Again, taking limits as t, x and y tend to zero gives:

      ux     u y  
t

x

y
 q   . (35)

This is the 2D mass conservation equation and it is still and exact equation in that no
approximations have yet been made. Clearly we can easily generate this to 3D by
simply adding in a z-flow term, uz, to give:

      ux     u y     uz  
t

x

y

z
 q .   (36)

As before, the pressure still does not yet appear in this equation. We must manipulate
the left side of Equation (35) (or 36) as before (Equation 14) to yield:

         P 
  . (37)
t P  t 

We then use Darcy’s Law on the right side of the equations. Assuming that z is in the
vertical direction, the Darcy velocities are given by:

k x  P  z 
ux       g  , (38a)
  x  x 

k y  P  z 
uy      g  , (38b)
  y  y 

©HERIOT-WATT UNIVERSITY December 2017 v2


19

k z  P  
uz       g , (38c)
  z  

where we note that the permeability may be anisotropic, i.e. k x  k y  k z . Using the
above equations for the Darcy velocity along with Equation (37) gives the 3D pressure
equation:

      P     k x  P z      k  P z  
     g    y    g 
P t x    x x   y    y y  

   kz  P 

z  

 z 
 
  g    q

(39)

Equation (39) is the 3D generalization of Equation 14, including gravity and also a well
source/sink term. Again, this is a non-linear partial differential equation (PDE) which
cannot generally be solved analytically.

3. The Two-Phase Flow Equations

3.1 Review of Two-Phase Flow Concepts


We now consider the equations which govern the flow of two phases through a porous
medium – e.g. oil-water, gas-oil air-water. In certain respects, the approach is very
similar to that used for single-phase flow in Section 2. We apply the mass conservation
equation to each of the phases separately. However, there is a little more “two-phase
physics” that we must be aware of before proceeding. For this reason, we first review
some of the key concepts of two-phase flow (See Topic 2).

Key Concepts: Ensure you are familiar with the following main ideas of two-phase flow
(where we assume the phases are oil and water):

 Phase Saturations, So and Sw: where So + Sw = 1.

 Formation volume factors, Bo and Bw (units (rb/stb) where, for example o =


osc/Bo, and osc is the oil density at standard conditions.

©HERIOT-WATT UNIVERSITY December 2017 v2


20

 The two-phase Darcy Law (with gravity) and relative permeabilities krw(Sw) and
kro(Sw).

kkro  Po z 
uo     o g 
o  x x 
kk  P z 
uw   rw  w   w g 
w  x x 

(Note that we’re back to considering flow in one direction here – the x-direction.)

 Phase pressures Po and Pw, and the concept of capillary pressure, Pc(Sw) = Po – Pw,
which acts as a constraint between the phase pressure difference at various
saturations, Sw.

3.2 Derivation of the Two-Phase Conservation Equations


As for the single-phase case, we will set up the material balance in a control volume as
shown in Figure 6 and Figure 7.

A useful quantity to define is the mass flux. Jo is the mass flux of oil and Jw the mass
flux of water. Mass flux is the mass flow rate per unit cross-sectional area. The
dimension are ML-2T-1, and possible units are kg.m-2s-1, or lbs.ft-2day-1.

Figure 6 and the paragraphs below show how the mass flux of oil may be calculated.
(The calculation for the water flux is similar.)

Figure 6
Definition of the oil flux, Jo.

©HERIOT-WATT UNIVERSITY December 2017 v2


21

(40)

 qo o  uo osc
Therefore, mass flow rate of oil per unit area =    uo o 
 A  Bo
(41)

But, by definition the above quantity is the flux, Jo (Figure 6) (where we note that
qo / A  uo is the Darcy velocity of oil).

J o  uo o ; J w  uw  w . (42)

However, usually the density is given at standard conditions (sc = standard conditions
= 60oF and 14.7 psia, or ~15.6oC and 1 bar). Then Bo, or Bw are used to convert the
densities to reservoir values.

uo osc uw  wsc
Jo  ; Jw  . (43)
Bo Bw

Now we used the flux expression to perform the mass balance on control block i in
Figure 7.

Figure 7
Control volume (block i) for two-phase flow.

©HERIOT-WATT UNIVERSITY December 2017 v2


22

Mass of oil flowing into block i over time step t

(44)

Mass of oil flowing out of block i over time t =  J o At i 1/2 (45)

Therefore change in mass of oil in block i over t


=   J o i 1/2   J o i 1/2  At . (46)

Remember that A = yz. Also, note that we changed the signs in Equation (46), so
that we have minus (the flow going out minus the flow going in), because we are going
on to form the differential equation.

The accumulation – i.e. change in the mass of oil in block i – is then calculated. Note
that in the two-phase case, we need to remember to take account of the saturation of
oil in the block. This is the proportion of the pore volume occupied by the oil.

Change of mass in block i over time (accumulation) = Mass in block at time t + t –


Mass in block at time t
  xyz So o t t   xyz So o t   So o t t   So o t  xyz .
(47)

Equating these two expressions, we get:


.
 So o t t   So o t  xyz    J o i 1/2   J o i 1/2  At (48)

We then divide throughout by xyzt to obtain:

 So o t t   So o t   J o i 1/2   J o i 1/2 


 . (49)
t x

Taking limits as x and t tend to zero, we obtain the differential form of the mass
conservation equation:
  S o  o  J
 o . (50)
t x

Using the fact that o  osc Bo , gives:

©HERIOT-WATT UNIVERSITY December 2017 v2


23

   So osc J o
  . (51)
t  Bo  x
u
Then substituting for J o  o osc gives:
Bo

   So osc    uo osc 
   . (52)
t  Bo  x  Bo 

Cancelling osc, which is a constant, gives:

   So    uo 
     . (53)
t  Bo  x  Bo 

Similarly, we can derive an equation for water:

   Sw    uw 
     . (54)
t  Bw  x  Bw 

The above equations are the exact differential forms of the oil and water mass
conservation equations. They involve no assumptions: they simply arise as a result of
the definition of various terms. However, these equations are not useful in this form
since they do not mention pressure, and this is what we measure in a reservoir – not
local (spatially distributed) velocities, uo and uw. We therefore need to use Darcy’s Law.

Substitution for uo and uw in Equations (53) and (54) gives:

   So    kkro  Po  z  
       o g   . (55)
t  Bo  x  o Bo  x  x  

   Sw    kkrw  Pw  z  
  
t  Bw  x  w Bw  x    w g x   . (56)
  

These equations express the mass conservation of oil and water, on the assumption
that the two-phase Darcy Law applies. They are therefore not exact (because Darcy’s
law is empirical).

©HERIOT-WATT UNIVERSITY December 2017 v2


24

At first glance, it appears that Equations (55) and (56) have 4 unknowns: Po, Pw, So and
Sw. Since there are only two equations, it seems impossible to solve. However, we
have two additional equations, which act as constraints:

Pc  Po  Pw and So  Sw  1 .

Therefore there are only two independent unknowns: e.g. Po and Sw.

Although we are only dealing with a 1D model here, the derivation of the two-phase
pressure equation becomes quite lengthy. At this point, we are going to simplify the
two-phase flow equations. The full equations are given at in an Appendix to Topic 4
(and are non-examinable).

3.3 The Simplified Two-Phase Pressure Equation


Let us make the following assumptions:
1. The viscosities of oil and water, o and w, are constant (x and P);
2. Both the rock and the fluids are incompressible ( constant and Bo = Bw = 1);
3. Capillary and gravity effects are negligible (Pc = 0; Po = Pw and z x  0 .

Simplifying Equations (55) and (56) gives:

So   kkro  P  
     , (57)
t x  o  x  

S w   kkrw  P  
     . (58)
t x  w  x  

So S w
Remember that So + Sw = 1. Therefore   0 . (If we increase the oil
t t
saturation by a certain amount, we must decrease the water saturation by the same
amount.) Now, if we add these two equations together, we get:

  kkro  P     kkrw  P  
        0 (59)
x  o  x   x  w  x  

which can be simplified to:

©HERIOT-WATT UNIVERSITY December 2017 v2


25

  P 
  o  w     0 (60)
x   x  

Where o and w are the phase mobilities given by:

kkro kkrw
o  and w  . (61)
o w

These phase mobilities are functions of the saturation through the relative
permeabilities, i.e. we have o(So) and w(Sw). To simplify further, we define a total
mobility as:

T  So   o  w (62)

Thus, the pressure equation becomes:

   P  
T    0 . (63)
x   x  

This is the simplified pressure equation. However, we also need to have another
equation to give the saturation. This 2nd equation can be given by either Equation (57)
or (58), e.g.

So   kkro  P  
     . (64)
t x  o  x  

Note that these two equations are coupled. That is, the coefficients in the pressure
equation, T depend on the saturation, and the saturation depends on the pressure
gradient P x . Therefore, if we solve the equations one at a time, we don’t know
the values to use for the coefficients. Here we consider how we could solve the
equations. In Topic 5 we go into a bit more detail.

In grid block, i, we have two unknowns, Poi and Soi. Consider the situation at time t,
which is the nth time step (t = n x t).

©HERIOT-WATT UNIVERSITY December 2017 v2


26

We wish to find Poi and Soi at the next time step, t = t+t, or n+1.

The problem is shown schematically in Figure 8.

Figure 8
Schematic diagram showing the update of the pressure and saturation in grid block, i,
over time step, t.

The pressure equation has terms which depend on So, so we need to know Soin 1 to
solve for the pressure. However, we can try the following strategy:

a) Use S oin , i.e. the known saturation values to calculate the coefficients in the
pressure equation (Equation 63).

b) Solve the pressure equations to get a first estimate of Poin 1

c) Now solve the saturation equations (Equation 64) using the latest values for
Poin 1 for the pressure-dependent flow terms, to get the Soin 1 values.

©HERIOT-WATT UNIVERSITY December 2017 v2


27

d) We now have values for Poin 1 and Soin 1 . However, we may calculate the values
more accurately by going back to step a) and using our new saturations, Soin 1
, to obtain a better solution for the Poin 1 values.

e) We could iterate through steps a) to d) until the process converges, i.e. the
newly calculated Soin 1 and Poin 1 do not change any more (or only change by a
tiny amount). We then accept these as our “accurate” new time level values,
and go on to the next time level.

The above description outlines a strategy for solving the equations. We shall deal with
the solution of equations in Topic 5.

3.4 The Full Two-Phase Pressure Equation


For completeness, we state the full pressure equation for two-phase flow here. The
full derivation is in the Appendix to Topic 4. The full pressure equation is:

(65)

where (So; Po) contains fluid and rock compressibility terms.

The full saturation equation is:

   So    k kro  P z  
     o g   (66)
t  Bo  x  o Bo  x x  

©HERIOT-WATT UNIVERSITY December 2017 v2


28

These two equations can be solved to estimate the evolution of the pressure and
saturation throughout a system with time. They can take account of compressibility of
both the fluid and the rock, and also they can take the effects of gravity and capillary
pressure into account.

4. Closing Remarks
In this Topic, we have derived the fundamental equations for single- and two-phase
flow in a porous medium. These equations are derived by applying:

MATERIAL BALANCE + DARCY’S LAW

In two phases, the mass conservation is applied to each phase. We have to remember
to include the phase saturation, the relative permeabilities and the capillary pressure.
(Topic 7 discusses relative permeability and capillary pressure in detail.)

For both single- and two-phase flow, the resulting equations for the compressibility
system are non-linear PDE’s. Moreover, the two-phase flow equations are coupled
equations. The equations cannot be solved analytically. The numerical solution of
these equations is discussed in Topic 5.

5. References
Carslaw, H.S. and Jaeger, J.C., 1959. “Conduction of Heat in Solids”, 2md Edition,
Oxford Clarendon Press.

Crank. J., 1975. “The Mathematics of Diffusion”, 2nd Edition, Oxford Clarendon Press.

6. Appendix to Topic 4: The Derivation of the Full Two-Phase


Pressure Equation
We start with Equations (55) and (56), which we’ll refer to as (A1) and (A2) here.

   So    kkro  Po  z  
       o g   . (A1)
t  Bo  x  o Bo  x  x  

   Sw    kkrw  Pw  z  
  
t  Bw  x  w Bw  x    w g x   . (A2)
  

©HERIOT-WATT UNIVERSITY December 2017 v2


29

Expand the left side of Equation (A1), using the product rule for differentiation and
then the chain rule:

   So    1    So  So   
    So       
t  Bo  t  Bo  Bo  t  Bo  t 

  1   Po    So  So     Po 


  So        
Po  Bo   t  Bo  t  Bo  Po   t 

  So     1  So      Po 
     So      
Bo  t   Po  Bo  Bo  Po    t 
(A3)

We wish to eliminate the underlined term.

Similarly, the left side of the water equation (A2) is:

   Sw    Sw     1  Sw      Po 
      Sw      
t  Bw  Bw  t   Po  Bw  Bw  Po    t 
(A4)

(You may notice that we’ve changed Pw to Po in Equation A4. Since we are just
interested in the variation certain quantities with pressure, the label does not matter.)

So Sw
We want to eliminate the time derivative. Since   0 , we can multiply (A3)
t t
by Bo  and (A4) by Bw  , and then add the two equations.

Bo    So   So     1  So      Po 
     Bo So       (A5)
 t  Bo   t   Po  Bo    Po    t 

Bw    Sw   S w     1  Sw      Po 
     Bw Sw       (A6)
 t  Bw   t   Po  Bw    Po    t 

Adding, we get:

©HERIOT-WATT UNIVERSITY December 2017 v2


30

   1    1 
 Bo So    Bw S w  
Bo    So  Bw    S w   Po  Bo  Po  Bw    Po 
      t 
 t  Bo   t  Bw   So    S w   
     
   Po    Po  
   1    1  1      Po 
  Bo So    Bw S w      
 Po  Bo  Po  Bw    Po    t 

(A7)

where we have used So + Sw = 1 on the right side of Equation (A7).

The terms in the square brackets in Equation (A7) are compressibility terms. We
simplify the equation, by lumping these together, i.e.:

   1    1  1    
  So ; Po    Bo So    Bw S w     (A8)
 Po  Bo  Po  Bw    Po  

The expression above (A7) was derived by multiplying the left sides of equations (A1)
and (A2) by Bo  and Bw  , respectively, and then adding them together. We now
need to do the same thing to the right side of these equations.

Po Bo   kkro  Po  z   Bw   kkrw  Pw  z  


  So ; Po     x   o g x     x   B  x    w g x  
t  x  o Bo     w w   

(A9)

We now use the capillary pressure, Pc(Sw) = Po – Pw, to eliminate Pw, i.e.

Pc  Sw  Po Pw P P P


  so w  o  c (A10)
x x x x x x

This gives:

©HERIOT-WATT UNIVERSITY December 2017 v2


31

Po Bo   kkro  Po  z  


  So ; Po        o g  
t  x  o Bo  x  x  
(A11)
B   kkrw  Po   Pc  z  
 w       w g  
 x   w Bw  x   x  x  

This is the 1D pressure equation for a two-phase compressible (fluids and rock) system.
Each of the terms is physically interpretable and we expand this out to see each of the
contributions more clearly. This is the same as Equation (65), shown in Section 3.4.

(A12)

The saturation equation (same as Equation (66) is:

   So    k kro  P z  
     o g   . (A13)
t  Bo  x  o Bo  x x  

©HERIOT-WATT UNIVERSITY December 2017 v2


32

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 5 – Solution of the Flow


Equations
Contents
1. Introduction ............................................................................................................ 3
2. Review of Finite Differences ................................................................................... 4
3. Application of Finite Differences to Partial Differential Equations ........................ 9
3.1 Explicit Finite Difference Approximation of the Linear Pressure Equation .......... 9
3.2 Implicit Finite Difference Approximation of the Linear Pressure Equation ....... 13
3.3 Implicit Finite Difference Approximation of the 2D Pressure Equation ............. 17
3.3.1 Discretisation of the 2D Pressure Equation................................................. 17
3.3.2 Numbering scheme for the 2D Pressure Equation ...................................... 19
3.3.3 Implicit Finite Difference Approximation of Non-Linear Pressure Equations
.............................................................................................................................. 20
4. Application of Finite Differences to Two-Phase Flow .......................................... 23
4.1 Discretisation of the Two-Phase pressure and Saturation Equations ................ 23
4.2 IMPES Strategy for Solving the Two-Phase Pressure and Saturation Equations 26
5. Numerical Solution of Linear Equations ............................................................... 27
5.1 Introduction to Linear Equations........................................................................ 27
5.2 General Methods for Solving Linear Equations .................................................. 29
5.3 Direct Method for Solving Linear Equations ...................................................... 29
5.4 Iterative Methods for Solving Linear Equations ................................................. 31
5.5 Comparison of Iterative and Direct Methods for Solving Numerical Equations 35
6. Direct Solution of the Non-Linear Equations of Multi-Phase Flow ...................... 36
6.1 Introduction to Sets of Non-Linear Equations .................................................... 36
6.2 Newton’s Method for Solving Sets of Non-Linear Equations ............................. 38
6.3 Newton’s Method Applied to the Non-Linear Equations of Two-Phase Flow ... 44
7. Closing Remarks.................................................................................................... 46
8. Appendix ............................................................................................................... 46

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 Calculate the forward, backward and central FD approximations for dP/dx.

 Calculate the FD approximation for d2P/dx2.

 Derive the explicit FD approximation of the linear pressure equation in 1D.

 Derive the implicit FD approximation of the linear pressure equation in 1D.

 Derive the FD approximation for the simplified two-phase pressure equation.

 Describe two types of method for solving linear pressure equations.


o Contrast the two methods.

 Apply a simple iterative method to solve a set of linear equations.

 Explain the Newton-Raphson method for solution of non-linear equations in 1D.

Note that once again, there are some lengthy derivations in this Topic. The student
should work through all of these in detail to understand each step of the process. All
the derivations are examinable, except the solution of the non-linear equation which
is given in the Appendix.

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
As noted in Topic 4, the multi-phase flow equations for real systems are so complex
that it is not possible to solve them analytically. In practice these equations can only
be solved numerically. There are several different ways in which the equations may be
discretised. Here we shall use the simplest approach, which is the finite difference
method.

After a brief review of finite differences, we go on to apply them to very simple systems
such as for the simplified 1D pressure equation (derived in Topic 4). This equation does
not need to be solved numerically but it demonstrates how explicit and implicit finite
difference solution can be developed. We then show how sets of linear equations arise
in solving implicit equations and we consider solution of the linear equations in an
elementary manner. We then consider how the 2D pressure equation is solved.

In the numerical solution of the multi-phase flow equations, we need to solve for
pressure and flow. An outline of how this is done is presented but we do not go into
great detail.

At the very start of this course, we considered a very simple “simulation model” for a
growing colony of bacteria. The number of bacteria, N, grew with a rate proportional
to N itself i.e. (dN/dt) = N, where  is a constant. We saw that this simple equation
t
had a well-known analytical solution, N  No e , where No is the number of bacteria
at time, t = 0. This exponential growth law provides the solution to our model.
However, we also introduced the idea of a numerical model, even although we could
solve the problem analytically. The numerical version of the model came up with an
algorithm (or recipe). In the Tutorial exercise in Topic 1, you should have compared
the results of the analytical and numerical models and found out that they get closer
as we take successively smaller time steps, t. In this case, we say that the numerical
model converges to the analytical model. In fact, in many areas of science and
engineering, we often apply a numerical technique to a problem we can already solve
analytically i.e. where we “know the answer”. Why would we do this? The answer is
that we might be testing the numerical method to see how closely it gives the right
answer. More commonly, we might test several - maybe 3 or 4 - numerical techniques
to determine which one works best. The phrase “works best” in the context of a
numerical method usually means gives the most accurate numerical agreement with
the analytical answer for the least amount of computational work. Note the
importance of this balance between accuracy and work for a numerical method. There
may be no point in having a numerical method that is “twice” as accurate (in some
sense) for ten times the amount of computational work.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

In Topic 4, we already met the equations for single-phase and two-phase flow of
compressible fluids through porous media. These turned out to be non-linear partial
differential equations (PDEs). Recall that a non-linear PDE is one where certain
coefficients in the equation depend on the answer we are trying to find e.g. Sw(x,t),
P(x,y,z,t) etc. For example, for single phase compressible flow, the equation for
pressure, P(x,t), in 1D is given by:

P   k  P 
c  P 
t x   x 
(1)

In this equation, both the generalised compressibility term, c(P) (of both the rock and
the fluid), and the density, (P), terms depend on the unknown pressure, P, which we
are trying to find. As noted previously, such non-linear PDEs are very difficult to solve
analytically and we must usually resort to numerical methods. The main topic of this
module is on how we solve the reservoir flow equations numerically. This process
involves the following steps:

(i) Firstly, we must take the PDE describing the flow process and “chop it up” into grid
blocks in space. This is known as spatial discretisation and, in this course, we will
exclusively use finite difference methods for this purpose.

(ii) When we apply finite differences, we usually end up with sets of non-linear
algebraic equations, which are still quite difficult to solve. In some cases, we do
solve these non-linear equations. However, we usually linearise these equations
in order to obtain a set of linear equations.

(iii) We then solve the resulting sets of linear equations. Many numerical options are
available for solving sets of linear equations and these will be discussed below. This
is often done iteratively by repeatedly solving them until the numerical solution
converges.

This module will deal successively with each of the parts of the numerical solution
process, (i) - (iii) above.

2. Review of Finite Differences


The main concept of finite difference approximation is best illustrated by the following
simple example where we refer to Figure 1. Study the notation in this figure, since it
is the basis of that used throughout this chapter. The main task of the finite difference
approach is to represent the derivatives of the function, P(x), in an approximate
manner, i.e.

©HERIOT-WATT UNIVERSITY December 2017 v2


5

dP P  2 P
, , , etc.
dx x x 2

First consider how we might approximate dP dx at xi using the quantities in Figure 1.


It is easy to see that there are three ways to do this.

Approximation 1 – Forward Difference (fd): take the slope between Pi and Pi+1 as
being approximately equal to  dP dx i at xi to obtain:

 dP  Pi 1  Pi
   . (2)
 dx i fd x

Approximation 2 – Backward Difference (bd): take the slope between Pi-1 and Pi as
being approximately equal to  dP dx i at xi to obtain:

 dP  Pi  Pi 1
   . (3)
 dx ibd x

Approximation 3 – Central Difference (cd): we could take the average of the forward
and backward difference approximations to give us  dP dx i at xi.

 dP  1  dP   dP   1  Pi 1  Pi Pi  Pi 1 
        
 dx i cd 2  dx i fd  dx ibd  2  x x 

 dP  Pi 1  Pi 1
   (4)
 dx i cd 2x

©HERIOT-WATT UNIVERSITY December 2017 v2


6

Figure 1
Notation for the application of finite difference methods for approximating
derivatives.

 dP 
Each of the above approximates to   as shown in Figure 2. But which is best?
 dx i
There is not an unqualified answer to this question, but let’s take a simple numerical
example, e.g. P(x) = ex, where we know the right answer. The values given in Figure 3
illustrate how the methods perform.

 dP 
We take xi = 1.0, and x = 0.1, and P 1     2.7183 .
 dx 1

©HERIOT-WATT UNIVERSITY December 2017 v2


7

Figure 2
The three types of finite difference approximations

 dP  3.0042  2.7183
    2.859  err  0.14
 dx i fd 0.1

 dP  2.7183  2.4596
    2.587  err  0.13
 dx ibd 0.1

 dP  3.0042  2.4596
    2.723  err  0.005 (5)
 dx i cd 0.2

Figure 3
A numerical example of the finite difference approaches. P(x) = ex.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

The quantity in square brackets after each of the finite difference approximations
above is the error i.e. the difference between that method and the right answer which
is 2.7183. As expected from the figure for this case, the forward difference answer is a
little too high (by ≈ +0.14) and the backward difference answer is a little too low (by ≈
-0.13). The central difference approximation is rather better than either of the previous
ones. In fact, we note that the fd and bd methods give an error of the order ∆x, the
grid spacing (where, as engineers, we are saying 0.14 ≈ 0.1!). The cd approximation, on
the other hand, has an error of order ∆x2 (where again we are saying 0.005 ≈ 0.1x0.1 =
0.01). More formally, we say that the error in the fd and bd approximations are "of
order Δx" which we denote, O(∆x), and the cd approximation has an error "of order
Δx2" which we denote, O(∆x2).

d 2P
Now consider the finite difference approximation of the second derivative, .
dx 2
Going back to Figure 1, the definition of the second derivative of P(x) at xi, is the change
in the slope dP dx at xi. Therefore, we can evaluate this derivative between xi-1 and
xi (i.e. the bd approximation) and then do the same between xi and xi+1 (i.e. the fd
approximation) and simply take the rate of change of these two quantities with respect
to x, as follows:

 dP   dP 
   
d 2P  dx i fd  dx ibd

dx 2 x

Pi 1  Pi Pi  Pi 1

  x x
x
Therefore
d 2 P Pi 1  Pi 1  2 Pi
 . (6)
dx 2 x 2

d 2P
The accuracy of may be estimated using the example shown in Figure 3. The
dx 2
correct answer is 2.7183. Equation (6) gives:

d 2 P 3.0042  2.4596  2  2.7183


  2.7200  err  0.0017 (7)
dx 2 0.12

©HERIOT-WATT UNIVERSITY December 2017 v2


9

Thus we can see that the error in this case is approximately 0.001, so is O(x2).

In the introductory section of Topic 1, we already applied the idea of finite differences
(although we did not call it that at the time) to the simple ordinary differential
equation:

dN
N
dt

Here, N (the number of bacteria in the colony) as a function of time, N(t), is the
unknown we want to find. Denoting Nn and Nn+1 as the size of the colony at times t
(labelled n) and t + t (labelled n+1), we applied finite differences to obtain:

dN N n1  N n
 Nn (8)
dt t

This gives:

N n1  1  t  N n .

This gave us a very simple algorithm for calculating Nn+1. We then took this as the
current value of N and applied the algorithm repeatedly.

3. Application of Finite Differences to Partial Differential


Equations

3.1 Explicit Finite Difference Approximation of the Linear Pressure


Equation
We have seen in Topic 4 that the flow equations are actually partial differential
equations (PDEs) since the unknowns, P(x,t) and Sw(x,t) depend on both space and
time. As an example of a linear PDE, we will take the simplified pressure equation
(Equations 26 and 27; Topic 4) as follows:

P k 2 P 2 P
  D (9)
t c x 2 x 2
h

©HERIOT-WATT UNIVERSITY December 2017 v2


10

where Dh is the hydraulic diffusivity. As we noted in Topic 4, this PDE is linear and has
known analytical solutions for various boundary conditions. However, we will neglect
these and apply numerical methods as an example of how to use finite differences to
solve PDEs numerically. To make things even simpler, we will take Dh = 1, giving the
equation:

P  2 P
 (10)
t x 2

This is the pressure equation for a 1D system where 0 < x < L, where L is the length of
the system. We can visualise this physically - much as we did in Topic 4, Section 2.1 -
using Figure 4. After the system is set constant at P = Po, the inlet pressure is raised
(at x = 0) instantly to P = Pin while the outlet pressure is held at Pout = Po.

These pressures, Pin and Pout, represent the constant pressure boundary conditions
(Mathematically these are sometimes called Dirichlet boundary conditions.) In other
words, these are set, as if by experiment, and the system between 0 < x < L must “sort
itself out” or "respond" by simply obeying Equation (10) above.

Figure 4
Diagram of the pressure propagation in a 1D (compressible) system.

We now approach this problem using finite differences as follows:


 Discretise the x-direction by dividing it into a numerical grid of size ∆x;
 Choose a time step, ∆t;
 Use the following notation:

©HERIOT-WATT UNIVERSITY December 2017 v2


11

Pi n current (known) values of P (for all i) at time level, n;

Pi n 1 new (unknown) P ( for all i, except at boundaries) at time level n+1.

 Fix the boundary conditions which, in this case, are as follows (Figure 4):

P1 = Pin and PNX = Pout = Po which are fixed for all t.

 Apply finite differences to Equation (10) using the above notation to obtain:

n 1
 P  Pi  Pi
n

   (11)
 t i t

  2 P  Pi ??1  Pi ??1  Pi ??
 2  (12)
 x i x 2

However, an issue arises in Equation (12) above as shown by the question marks on
the spatial derivative time levels. It is simply: which time level should we take for the
spatial derivative terms in Equation (12)? This is important and we will return to this
matter soon. However, for the moment, let us take these spatial derivatives at time
level n (the “known” level) since this will turn out to be the simplest thing we can do.
Thus we obtain:

  2 P  Pi n1  Pi n1  2 Pi n
 2  (13)
 x i x 2

Substituting these approximations into Equation (10) gives:

Pi n1  Pi n Pi n1  Pi n1  2Pi n


 (14)
t x 2

n 1
which arranges into an explicit expression for Pi , the only unknown in the above
equation:

t
Pi n 1  Pi n 
x 2
 Pin1  Pin1  2Pi n  (15)

In other words, we have an algorithm which says:

©HERIOT-WATT UNIVERSITY December 2017 v2


12

New value (level n+1) of Pi = Old value (level n) of Pi + “a correction term”.

Equation (15) gives the algorithm for propagating the solution of the PDE forward in
time from the given set of initial conditions.

Let us consider an example with initial conditions and boundary conditions:

P1  1 for all time (boundary condition)

Pi 0  0 at time t = 0, for 2 < i < NX – 1

PNX  Po for all time (boundary condition)

In this example, assume that:

0  x  1.0

x  0.1  implies 11 grid points, P1, P2, P3, ..., P11 (NX = 11)

t  0.01 .

The problem is then to calculate the solution, P(x,t), that is for all i (at all grid points)
and all future times up to some final time (possibly when the equation comes to a
steady-state as will happen in this example). This exercise is carried out in Tutorial 1
for Topic 5.

The assumption we made in Equation (12) was that the spatial derivative was taken at
the n (known) time level. This allowed us to develop an explicit formula for the
pressure (for all i). This method is therefore known as an explicit finite difference
method. We can learn about some interesting and useful features of the explicit
method which is encapsulated in Equation (15) by simply experimenting with the
spreadsheet in Tutorial 1 for this Topic. Three features of this method can be
illustrated by “numerical experiment” as follows:

(i) The effect of time step size, t;

(ii) The effect of refining the spatial grid size, x (or number of grid cells, NX);

©HERIOT-WATT UNIVERSITY December 2017 v2


13

(iii) The effect of running the calculation to steady-state as t   (in practice, until
the numerically computed solution stops changing).

Another way to represent this explicit finite difference solution to the PDE is shown in
Figure 5, where we indicate the time levels for the solution of the equations and we
n 1
also show the dependency of the unknown pressure ( Pi ).

Figure 5
Schematic diagram of the explicit finite difference algorithm for solving the linear
pressure diffusion equation.

3.2 Implicit Finite Difference Approximation of the Linear Pressure


Equation
We now return to the original finite difference Equation (12), where we had to make a
choice of time level for the spatial derivative,  P  x . We now examine the
2 2

consequences of taking this derivative at the (n+1) time level - this is the “unknown”
time level. The finite difference equation for this case is as follows: The time derivative
is the same, i.e.
n 1
 P  Pi  Pi
n

   (16)
 t i t

but the spatial derivative is now:

  2 P  Pi n11  Pi n11  2 Pi n 1
 2  . (17)
 x i x 2

As before, we equate the finite difference approximations to obtain:

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Pi n 1  Pi n Pi n11  Pi n11  2 Pi n 1
 . (18)
t x 2

This equation should be compared with Equation (14). In the previous case, the
n 1
equation could be easily rearranged into an explicit equation for Pi (Equation 15).
However, Equation (18) cannot be rearranged so easily. There are now 3 unknowns.
This seems to be a paradox: how do we calculate 3 unknowns from one equation. The
answer is that there is one such equation for each grid point. We show below that this
is a set of linear equations, and that there are the same number of equations as there
are unknowns. This approach is more work than the simple explicit method. We have
an implicit set of equations, and this method is known as an implicit finite difference
method.

To obtain the linear equations, rearrange Equation (18) so that all of the unknowns are
on the left and the knowns are on the right.

 x 2  n 1  x 2  n
Pi n11   2  n 1
 Pi  Pi 1     Pi . (19)
 t   t 

This equation has the form:

ai 1Pi n11  ai Pi n1  ai 1Pi n11  bi (20)

 x 2   x 2  n
where ai 1  1; ai    2   i 1
; a  1 and b     Pi .
t   t 
i

These are all known. The ai do not change through the calculation, but bi is updated at
each time step as the newly calculated Pn+1 is set to Pn for the next time step. This is
demonstrated below using a system with 5 grid blocks as shown in Figure 6. We
assume that the pressures are fixed on the boundaries: P1 at grid point 1 and P5 at
grid point 5.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

Figure 6
Simple example of a 5 grid-block system to illustrate the implicit finite difference
scheme.

We write out the equations for points 2, 3 and 4, which we wish to solve for all times
after t = 0. (We can assume that the pressures are all equal to P5 at time t = 0.)

1) P1 = P1 ;

 x 2  n
2) P1  a2 P2n 1  P3n 1  b2     P2
 t 

 x 2  n
 a2 P2n 1  P3n 1     P2  P1
  t 

 x 2  n
3) P2n 1  a3 P3n 1  P4n 1  b3     P3
 t 

 x 2  n
4) P3n 1  a4 P4n 1  P5  b4     P4
 t 

 x 2  n
 P3n 1  a4 P4n 1     P4  P 5
 t 

5) P5 = P5 . (21)

As you can see, we have three equations for three unknowns. We can represent these
equations in matrix form, as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


16

  x 2  n 
  P2  P1 
  t  
 a2 1 0   P2n 1  
1    x  n 
2

 a3 1   P3n 1       P3  (22)
   t  
 0 1 a4   P4n 1   
   x  P n  P 
2

  t  4 5
 
The structure of this matrix equation is clearer when there are more equations
involved. For example, it is quite easy to show that, if we take 12 grid points instead
of the 5 above, we obtain 10 equations (using the two fixed boundary conditions,
P1  P1 and P12  P12 ) for the quantities P2n1 , P3n1 , P4n1 ,..., P11n1 of the form:

(23)

For this simple 1D PDE, it is clear that the matrix arising from our implicit finite
difference method has the following properties:

(i) It is tridiagonal - that is, it has a maximum of three non-zero elements in any
row and these are symmetric around the central diagonal;

(ii) It is very sparse - that is, most of the elements are zero. In an MxM matrix,
there are only 3M non-zero terms but M2 actual elements. If M = 100, then
the matrix is only (300/1002)x100% = 3% filled with non-zero terms.

Finally, note that the implicit finite difference scheme may be viewed as shown in
Figure 7. Contrast this with the equivalent figure for the explicit method (Figure 5).

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Figure 7
Schematic diagram of the implicit finite difference algorithm for solving the linear
pressure diffusion equation.

3.3 Implicit Finite Difference Approximation of the 2D Pressure Equation


Before going on to derive the finite difference equations for a two-phase system, we
will consider an extension of the equations in to 2D. We leave the extension of the
equations of the explicit method to 2D as an exercise for the student. Here we describe
the extension of the implicit equations to 2D and we allow the system to be anisotropic
(the permeability is different in different directions, i.e. k x  k y ).

3.3.1 Discretisation of the 2D Pressure Equation


We start with the basic pressure equation for single-phase, slightly compressible flow.
Equation (39) of Topic 4 gives the 3D generalisation of the pressure equation. If we
simplify this equation by considering only 2D (x-y plane), and ignoring gravity and
source/sink terms, we have:

     P     k x  P      k y  P  
        . (24)
P  t  x    x   y    y  

Making similar assumptions to those in Section 2.3 in Topic 4, we get:

P 2 P 2 P
 Dh, x 2  Dh, y 2 (25)
t x y

kx ky
where Dh , x  and Dh , y  is the hydraulic diffusivity in the x- and y-
 c  c
directions, respectively.

Equation (25) above can be discretised in a similar way to that applied to the 1D linear
pressure PDE discussed above. However, we will need to be quite clear about our

©HERIOT-WATT UNIVERSITY December 2017 v2


18

notation in 2D and, for this purpose, we refer to Figure 8. This shows the discretisation
grid for the above PDE - note that this is essentially the opposite of what we did when
we derived the equation in the first place! In Topic 4, we used a control volume (or
grid block) to express the mass conservation and then inserted Darcy’s Law for the
block to block flows; we then took limits as x, y and t 0. Here, we are starting
with the PDE and going back to the local conservation of flows and introducing finite
size x and y.

Figure 8
Discretisation and notation for the 2D pressure equation.

Discretising equation (25) above and using the notation shown in Figure 8, we get:

Pi ,nj1  Pi ,nj
 Dh, x
P n 1
i 1, j  Pi n1,1j  2 Pi ,nj1 
 Dh, y
P n 1
i , j 1  Pi ,nj11  2 Pi ,nj1 
(26)
t  x   y 
2 2

Note that in this case, we’ve not assumed that the hydraulic diffusivity equals 1, like
we did in Equation (10). Since we are using the implicit scheme, we take the pressure
on the right side at time (n+1).

Collecting terms in (n+1) on the left side, we get:

Dh , x t Dh , x t Dh , x t Dh , y t
Pi n1,1j  Pi n1,1j  2 Pi ,nj1  Pi ,nj11
x 2
x 2
x 2
y 2

(27)
Dh , y t n 1
Dh , y t n 1 n 1
 P i , j 1 2 P P  P n

y 2 y 2
i, j i, j i, j

©HERIOT-WATT UNIVERSITY December 2017 v2


19

This is an equation of the form:

ai , j 1Pi ,nj11  ai 1, j Pi n1,1j  ai , j Pi ,nj1  ai 1, j Pi n1,1j  ai , j 1Pi ,nj11  bi , j (28)

where the coefficients ai,j and bi,j are known. Note that the terms have been arranged
in the order south, west, centre, east and north for convenience later (Figure 8).

Equation (28) is similar to Equation (20) for the 1D system. However, now we have 5
non-zero terms in each equation rather than 3, and the matrix of coefficients is known
as a pentadiagonal matrix.

3.3.2 Numbering scheme for the 2D Pressure Equation


It is convenient to label the pressures as Pi,j when we are working out the discretisation
of the equations, but this is not helpful when we are arranging the linear equations.
Here, it is useful first to consider the numbering scheme for the 2D system that allows
us to dispense with the (i,j) subscripting in Equations 27 or 28 above. The structure of
the A-matrix in Equation (28) (with elements aij) can be made clearer by working out a
specific 2D example as shown in Figure 9.

Figure 9
Numbering system for the 2D equations.

Using the index, m, the equations become:

amnx Pmnnx
1
 am1Pmn11  am Pmn1  am1Pmn11  amnx Pmnnx
1
 bm (29)

Note that, when we apply the above equation numbering scheme to the example in
Figure 9 (nx = 4, ny = 5 and therefore, 1  m  20 ), certain “neighbours” are “missing”
when a block is at the boundary (or in a corner where two neighbours are missing).

©HERIOT-WATT UNIVERSITY December 2017 v2


20

For example, for block (i = 1; j = 3), that is block m = 9, the “i-1 block” is not there.
Therefore, the coefficient am-1 =0 in this case. This is best seen by writing out the
structure of the 20 x 20 A-matrix by referring to Figure 9; the A-matrix structure is
shown in Figure 10.

Note that the A-matrix structure in Figure 10 is sparse and has a maximum of five non-
zero coefficients in a given row – it is a pentadiagonal matrix.

Figure 10
Structure of the matrix of coefficients for the 2D pressure equation. See Figure 9 for
the ordering scheme for the cell counter, m.

All implicit methods for discretising the pressure equation lead to sets of linear
equations. These have the general matrix form:

A.x  b (30)

where A is a matrix like the examples shown above, x is the column vector of unknowns
(such as pressures) and b is a column vector of the RHSs. This is just like Equation (31)
but it is in shorthand form. We will discuss methods for solving these equations later
in this chapter. For the meantime, we will just assume that it can be solved. We next
consider when the PDEs describing a phenomenon are non-linear PDEs.

3.3.3 Implicit Finite Difference Approximation of Non-Linear Pressure Equations


In Sections 3.3.1 and 3.3.2, we have assumed that the equations are linear – i.e. the
coefficients do not depend on pressure. However, if we are dealing with compressible

©HERIOT-WATT UNIVERSITY December 2017 v2


21

systems (i.e. compressible rock and fluid) the equations are non-linear. For example,
the density of the rock and fluid may depend on the pressure, which is what we’re
trying to solve. We will not derive the non-linear equations here, but the method is
similar to the method for deriving the linear equations. The final result is of the form:

 mn1nx Pmnnx1  mn11Pmn11  mn1Pmn1  mn11Pmn11  mn1nx Pmnnx1  mn1 (31)

where we have used Greek symbols to indicate that the coefficients depend on
pressure. The elements of the matrix, m and the column vector  are not constant,
because they depend on pressure.

How do we go about solving the non-linear set of algebraic equations in (31)? We have
two choices in tackling this more difficult problem as follows:

(i) We can use a numerical equation solver which is specifically designed to solve
more difficult non-linear problems. An example of this type of approach is in using
the Newton-Raphson method. We will return to this method later (once we have
seen how to solve linear equations).

(ii) We can choose to apply a more pragmatic algorithm such as the following:

a) Although our -terms in Equation (31) are strictly at time level (n+1), simply
take them at time level n, as a first guess. So we approximate Equation (31) by
the following first guess:

 mn nx Pmnnx1  mn 1Pmn11  mn Pmn1  mn 1Pmn11  mn nx Pmnnx1  mn (32)

where the quantities m, etc and  m are evaluated at the (known) time level n,
n
i.e. at values of pressure, Pm .

b) Solve the now linear Equations (32) above to obtain a first estimate - or a first
iteration - of the Pn each grid point. We will use the following notation:  m


where  is the iteration counter and  m is the value of the  coefficient for

P 
m
n 1 
, after  iterations; that is:

 m   m  Pmn1   .

(33)
 

©HERIOT-WATT UNIVERSITY December 2017 v2


22

c) Use the latest iterated values of the to solve the (now linear) equations to go
from to Pm  to  P 
n 1 v n 1 v 1
m .

d) Keep iterating the above scheme until it converges; i.e. the difference between
the sum of two successive iterated values of the pressure (Err) is sufficiently
small (< Tol, which is an acceptably small value):

Err    Pmn1    Pmn1 


 1 
(34)
all m

Stop if Err < Tol. Otherwise continue through steps above.

The algorithm outlined in (ii) above for solving the non-linear pressure equation is
represented in Figure 11.

Figure 11
Algorithm for the numerical solution of the non-linear 2D pressure equation.

©HERIOT-WATT UNIVERSITY December 2017 v2


23

4. Application of Finite Differences to Two-Phase Flow

4.1 Discretisation of the Two-Phase pressure and Saturation Equations


We now consider how finite difference methods are applied to the two-phase flow
equations. We have seen how these equations were derived in Topic 4. Recall that, in
two-phase flow, we have two coupled equations to solve - a pressure equation and a
saturation equation. For example, we may solve for the oil pressure, Po(x,t) and the oil
saturation, So(x,t); we would then find the water saturation and water pressure by
using the constraint equations, So + Sw = 1 and the capillary pressure relation, Pc(Sw) =
Po - Pw, respectively.

From Topic 4, we use the highly simplified form of the 1D pressure and saturation
equations, where we choose to solve for P(x,t) and So(x,t) as shown in Topic 4,
Equations (63) and (64). Note that we have taken zero capillary pressure (therefore P
= Po = Pw) and a horizontal system, so no gravity, which gives:

   P  
PRESSURE EQUATION  T  So      0 (35)
x   x  

 So     P 
SATURATION EQUATION     o  So     (36)
 t  x   x  

(Equations (35) and (36) are the same as Equations 63 and 64 from Topic 4,
respectively).

The quantity T is the total mobility and is the sum of the oil and water mobilities:

T  So   o  So   w  So  , where  f  kkrf  f and f = o or w.

Equations (35) and (36) are clearly coupled together since the oil saturation appears in
the non-linear coefficient of the pressure equation. Likewise, the pressure appears in
the flow term on the right side of the saturation equation.

We can now apply finite differences to each of these equations in the same way as
discussed above to obtain the pressure equation:

©HERIOT-WATT UNIVERSITY December 2017 v2


24

  P     P 
T  So   x    T  So    
    i 1/2   x   i 1/2
0 (37)
x

which can be further simplified to give:

  Pi n11  Pi n 1     Pi n 1  Pi n11  
 
 T o 
S     
 T o 
S  
  x   i 1/2   x   i 1/2
0 (38)
x

In Equation (38), we must specify the time level of the non-linear mobility. We could
either take this at time level n or (n+1). We will choose the most difficult option, which
is to specify at the new time, i.e. at Son 1 . There terms are denoted at T So
n 1
  i 1/2


and T So
n 1
 i 1/2
. This gives:


 T  Son1   
  T  Son 1   

 x 2
i 1/2
P n 1
i 1  Pi n 1
   x2 i 1/2
 i i1   0 (39)
P n 1
 P n 1 

   

Rearranging, we obtain:

  S 
T
n 1
o
i 1/2
P n 1


 T  Son 1  
i 1/2

  S 
T
n 1
o
i 1/2

 Pi 
n 1

T  Son 1  
i 1/2
Pi n11  0
i 1
x 2
 x 2 x 2
 x 2
 
(40)

This is a non-linear set of algebraic equations for the unknown pressures,


Pi n11 , Pi n1 and Pi n11 , but the coefficients depend on the – also unknown – saturations,
Son 1 . Note that Equation (40) represents one of a set of equations since there is one
at each grid point (and at the ends of the 1D system the values may be set by the
boundary conditions).

We now consider the discretisation of the saturation Equation (36). Finite differences
may be applied as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


25

  P     P  
 o  So      o  So    
S S    x   i 1/2   x   i 1/2
n 1 n
 
oi oi
(41)
 t  x

Again we can expand the derivatives at the (i+1/2) and (i-1/2) boundaries (Figure 8)
and take the mobility terms at time level (n+1), to obtain:

 Soin 1  Soin   o  So i 1/2 n1   o  Son 1  


n 1



 
  2  Pi 1  Pi
n 1
 

 
  2
i 1/2
 Pi
n 1
 Pi
n 1
1  

 t   x
 
x
(42)

The above non-linear set of equations can be written in various ways. Two particularly
useful ways to write the above equations are as follows:

Form A:

  o  So i 1/2 n 1
t    o  Son 1  
n 1

S n 1
oi  S  
n

 
oi
x 2  Pi 1  Pi
n 1
   
  x 2
i 1/2
 Pi
n 1
 P n 1
i 1  
 
 
(43)

Form B:

  o  So i 1/2 n 1
t    o  Son 1  
n 1

S n 1
oi S  
n

 
oi

x 2  Pi 1  Pi  
n 1
 
  x 2
i 1/2
 Pi  Pi 1     0
n 1 n 1

 
 
(44)

The reason for writing the above two forms of the saturation equation is that each is
useful, depending on how we intend to approach the solution of the coupled pressure
and saturation equations (Equations 63 and 64).

As with the case of the solution of the non-linear single phase pressure equation for a
compressible system, we have two strategies which we can use to solve the two-phase
equations above, as follows:

We can actually use a numerical equation solver which is specifically designed to solve
more difficult non-linear problems; e.g. the Newton-Raphson method. We will return
to this method later (once we have seen how to solve linear equations).

©HERIOT-WATT UNIVERSITY December 2017 v2


26

We can again choose to apply a more pragmatic algorithm and this is the subject of
Section 4.2 below.

4.2 IMPES Strategy for Solving the Two-Phase Pressure and Saturation
Equations
Take the pressure and saturation equations as follows (Equations 40 and 43):

PRESSURE:

  S 
T
n 1
o
i 1/2
P n 1

 T  Son 1 


i 1/2

  S 
T
n 1
o
i 1/2
 
  S n 1 
 Pi n 1  T o

i 1/2
Pi n11  0
i 1
x 2
 x 2 x 2
 x 2
 
(40)

SATURATION:

  o  So i 1/2 n 1
t    o  Son 1  
n 1

S n 1
oi  S  
 
n
oi
x 2  Pi 1  Pi
n 1
   
  x 2
i 1/2
 Pi
n 1
 Pi 1  
n 1

 
 
(43)

(i) In order to be able to solve the pressure equation, take the coefficients at time
level n (i.e. use the mobilities for the known saturations, S on ). This gives a first
estimate for the pressures.
(ii) Use the first estimate of the pressures to update the saturations, using Equation
(43). This gives us an estimate of Son 1 .
(iii) Use the updated saturation values to give another estimate for Pn+1.
(iv) Iterate these steps, using  as an iteration counter. Keep iterating these steps until
P  n 1  1
  Pi n 1 

the solution converges – i.e. until i
is less than some small

tolerance (e.g. 10-6). Now we have solved the pressure equations implicitly.
(v) Then the saturations may be solved explicitly using Equation (43).

This approach is known as the IMPES approximation, which stands for IMplicit Pressure
Explicit Saturation. There are however limitations in the time steps, t in the IMPES
approach. If t is too large, the solution may become unstable and give unphysical
results.

©HERIOT-WATT UNIVERSITY December 2017 v2


27

5. Numerical Solution of Linear Equations

5.1 Introduction to Linear Equations


In all implicit finite difference methods, we end up with a set of linear equations to
solve. In fact, we have shown above that the equations involved - e.g. the discretised
pressure equation - may be a set of non-linear algebraic equations. However, these
can usually be linearised (say, by time lagging the coefficients as discussed above) in
order to obtain a set of linear equations. The solution to the original non-linear
equations may then be found by repeating or iterating through the cycle of solving the
linear equations. In the following section, we will address the issue of solving the non-
linear equations which arise in the discretised two-phase flow equations directly.

A generalised set of linear equations may be written in full as follows:

a11 x1  a12 x2  a13 x3  a14 x4  a15 x5 ...  a1n xn  b1

a21 x1  a22 x2  a23 x3  a24 x4  a25 x5 ...  a2n xn  b2

a31 x1  a32 x2  a33 x3  a34 x4  a35 x5 ...  a3n xn  b3

….. …..

….. …..

….. …..

an1 x1  an 2 x2  an3 x3  an 4 x4  an5 x5 ...  ann xn  bn


(45)

Where the aij and bi are known constants and the xi (i = 1, 2, …, n) are the unknown
quantities which we are trying to find. The xi in our pressure equation, for example,
would be the unknown pressures at the next time step.

The set of linear equations can be written in matrix form as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


28

 a11 a12 a13 a14 a15 ... a1n   x1   b1 


a a22 a23 a24 a25 ... a2 n   x2  b2 
 21
 a31 a32 a33 a34 a35 ... a3n   x3   b3 
    
 ... ... ... ... ... ... ...   .    . 
 ... ... ... ... ... ... ...   .   . 
    
 ... ... ... ... ... ... ...   .   . 
a ann   xn  bn 
 n1 an 2 an 3 an 4 an 5 ...
(46)

We can write the matrix A and the column vectors x and b as follows:

 a11 a12 a13 a14 a15 ... a1n   x1   b1 


a a22 a23 a24 a25 ... 
a2 n    b 
 21  x2   2
 a31 a32 a33 a34 a35 ... a3n   x3   b3 
     
A   ... ... ... ... ... ... ...  ; x   .  ; b   . 
 ... ... ... ... ... ... ...  . .
     
 ... ... ... ... ... ... ...  . .
a    b 
 n1 an 2 an 3 an 4 an 5 ... ann   xn   n
(47)

and the set of linear equations can be written in a very compact form:

A.x  b (48)

where A is an n x n square matrix and x and b are n x 1 column vectors.

A simple equation of a set of linear equations is given below:

4 x1  2 x2  1x3  13
1x1  3x2  3x3  14 (49)
2 x1  1x2  2 x3  11

How do we solve these? In the above case, you can do a simple trial and error solution
to find that x1 = 2, x2 = 1 and x3 = 3, although this would be virtually impossible if there
were hundreds of equations. Remember, there is one equation for every grid block in
a linearised pressure equation in reservoir simulation (although there are lots of zeros

©HERIOT-WATT UNIVERSITY December 2017 v2


29

in the A matrix). This implies that we need a clear numerical algorithm that a computer
can work through and solve the linear equations. The overview of approaches to
solving these equations is discussed in the next section.

5.2 General Methods for Solving Linear Equations


Firstly, let us note that there is a vast literature on solving sets of linear equations and
many books on the underlying theory and on the numerical techniques have appeared.
Indeed, petroleum reservoir simulation has led to the development of some of these
numerical techniques. We will not cover much of this huge subject but we will cover
enough that the student can appreciate - rather than understand in detail - how the
linear equation are solved in a simulator.

Basically, there are two main approaches for solving sets of linear equations involving
direct methods and iterative methods as follows:

Direct Methods: this group of methods involves following a specific algorithm and
taking a fixed number of steps to get to the answer (i.e. the numerical values of x1, x2
.. xn). Usually, this involves a set of forward elimination steps in order to get the
equations into a particularly suitable form for solution (see below), followed by a back
substitution set of steps which give us the answer. An example of such a direct method
is Gaussian Elimination.

Iterative Methods: in this type of method, we usually start off with a first guess (or
estimate) to the solution vector, say x(0) . Where we take this as iteration zero,  = 0.
We then have a procedure – an algorithm – for successively improving this guess by
( )
x
 x ....  x etc. If it is successful, then this
(1) (2) (3)
iteration to obtain, x
iterative method should converge to the correct x as  increases. The solution should
get closer and closer to the “correct” answer but, in many cases, we cannot say exactly
how quickly it will get there. Therefore, iterative methods do not have a fixed number
of steps in them as do direct methods. Examples of iterative methods are the Jacobi
iteration, the LSOR (Line Successive Over Relaxation) method, etc.

The following two sections will discuss each of these approaches for solving sets of
linear equations in turn.

5.3 Direct Method for Solving Linear Equations


In fact, we will not present the details of any direct method for solving linear equations.
Instead, we will discuss a general outline of how these methods work. Starting with
the basic linear equation in matrix form:

©HERIOT-WATT UNIVERSITY December 2017 v2


30

A.x  b (50)

where A is an n x n square matrix and x and b are n x 1 column vectors. We can form
an augmented matrix with (n+1) columns and n rows, as follows:

 a11 a12 a13 a14 a15 ... a1n b1 


a a22 a23 a24 a25 ... a2 n b2 
 21
 a31 a32 a33 a34 a35 ... a3n b3 
 
 ... ... ... ... ... ... ... .  (51)
 ... ... ... ... ... ... ... . 
 
 ... ... ... ... ... ... ... . 
a ann bn 
 n1 an 2 an 3 an 4 an 5 ...

Remember that any mathematical operation we perform on one of the linear


equations (e.g. multiplying through by 2) must be performed on both sides of the
equation i.e. on the aij and the bi. Therefore, suppose we can transform the above
augmented matrix into the following form (we do not say how this is done, just that it
can be done):

c11 c12 c13 c14 c15 ... c1n e1 


0 c22 c23 c24 c25 ... c2 n e2 

0 0 c33 c34 c35 ... c3n e3 
 
 ... ... ... ... ... ... ... .  (52)
 ... ... ... ... ... ... ... . 
 
 ... ... ... ... ... ... ... . 
0 cnn en 
 0 0 0 0 ...

The C-matrix above is in upper triangular form, i.e. only the diagonals and above are
non-zero
( cij  0, j  i ) and all elements below the diagonal are zero ( cij  0, i  j ), as shown
above. Now consider why this particular form is of interest in solving the original
equations. The reason is that, if we have formed the augmented matrix correctly (i.e.
doing the same operations to the A and b coefficients), then the following matrix
equation is equivalent to the original matrix equation:

©HERIOT-WATT UNIVERSITY December 2017 v2


31

c11 c12 c13 c14 c15 ... c1n   x1   e1 


0 c22 c23 c24 c25 ... c2 n   x2  e2 

0 0 c33 c34 c35 ... c3n   x3   e3 
    
 ... ... ... ... ... ... ...   .    .  (54)
 ... ... ... ... ... ... ...   .   . 
    
 ... ... ... ... ... ... ...   .   . 
0 cnn   xn  en 
 0 0 0 0 ...

This matrix equation is very easy to solve, as can be seen by writing it out in full as
follows:

c11 x1  c12 x2  c13 x3  c14 x4  c15 x5 ...  c1n xn  e1

c22 x2  c23 x3  c24 x4  c25 x5 ...  c2n xn  e2

c33 x3  c34 x4  c35 x5 ...  c3n xn  e3

c44 x4  c45 x5 ...  c4n xn  e4


….. …..

….. …..

cn1,n1 xn1  cn1,n xn  en1

cnn xn  en (55)

We can now easily solve this equation by back-substitution, starting from equation n
in which we only have one term to find, xn. Once xn has been determined, it can be
used in the (n-1) equation to determine xn-1, etc.

5.4 Iterative Methods for Solving Linear Equations


As noted above, the idea in an iterative method for solving a set of linear equations is
to make a first guess at the solution and then to refine it in a stepwise manner using a
suitable algorithm. This procedure should gradually converge to the correct answer.
We illustrate the idea of such a method using a very simple iterative scheme. Note
that this simple point iterative scheme will work for some of the examples we try here
but it is not one that is recommended for use in reservoir simulation. However, it

©HERIOT-WATT UNIVERSITY December 2017 v2


32

adequately illustrates the main ideas which you need to know for the purposes of this
part of the course.

Our simple scheme starts with the longhand version of a set of linear equations and,
for our purposes, we will just take a set of four linear equations as follows:

a11 x1  a12 x2  a13 x3  a14 x4  b1


a21 x1  a22 x2  a23 x3  a24 x4  b2
(56)
a31 x1  a32 x2  a33 x3  a34 x4  b3
a41 x1  a42 x2  a43 x3  a44 x4  b4

Rearrange the equations as follows:

1
x1  b1   a12 x2  a13 x3  a14 x4  
a11 
1
x2  b2   a21 x1  a23 x3  a24 x4  
a22 
(57)
1
x3  b3   a31 x1  a32 x2  a34 x4  
a33 
1
x4  b4   a41 x1  a42 x2  a43 x3  
a44 

The resulting equations are precisely equivalent to the original set. They have been
reorganised to form the basis for an iterative scheme. Firstly, we introduce the
following notation:

xi( ) - the solution for xi at iteration , where  is the iteration counter;

( )
x , x - the first guess and the th iteration of the solution vector, x.
(0)

Err - an estimate of the error in the iterative scheme from the th to the (-1)th
iteration

Tol - some “small” quantity which determines the acceptable error in the iteration
scheme.
If Err < Tol the scheme has converged.

©HERIOT-WATT UNIVERSITY December 2017 v2


33

Using the above notation in Equations (57) above, we get the following simple iterative
scheme.

x1 1 
1 
a11   
b1  a12 x2 v   a13 x3   a14 x4  

x2 1 
1 
a22 

b2  a21 x1 v   a23 x3   a24 x4  
 
(58)
x3 1

1 
a33 

b3  a31 x1 v   a32 x2   a34 x4  

x4
 1

1 
a44  
b4  a41 x1   a42 x2   a43 x3  
v  
 
This iteration scheme may now be applied as follows:

(i) Make an initial guess at the solution, iteration



  0 : x 0  x1 0 , x2 0 , x30 , x40 
(ii) Update the solution to the next iteration  + 1 using Equations 58;

(iii) Estimate the error term (Err) by comparing the ( + 1) values with the  values:

4
Err   xi 1  xi
i 1

(iv) If Err < Tol, the scheme has converged. If not go back to step (ii) and continue
the iterations.

This scheme is illustrated below with a practical example.

Example: Solve the following set of equations using the iterative scheme described
above.

3.1x1  0.32 x2  0.5 x3  0.1x4  13.92

0.2 x1  2.1x2  0.33 x3  0.21x4  5.63

0.23 x1  0.32 x2  4.0 x3  0.3 x4  15.19

0.42 x1  0.22 x2  0.5 x3  5.2 x4  16.76

©HERIOT-WATT UNIVERSITY December 2017 v2


34

Take the first guess as: x1   x2   x3   x4   2 .


0 0 0 0

Hint: re-organise the above equations as follows:

x1
 1
 
 1 3.1  13.92  0.32 x2   0.5 x3   0.1x4
  

x2  1 2.1   5.63   0.20 x  033x  0.21x 
 1      
1 3 4

x3 1  1 4.0   15.19   0.23x  0.32 x  3.0 x


 
1 2
   
4 
x4  1 5.2   16.76   0.42 x  0.22 x  0.5 x 
 1      
1 2 3

Some terms are underlined to compare with those underlined below.

Fill out the following table using Excel or Matlab. (This may be tried in the Tutorial 2
for Topic 5.)

Iteration
counter x at iteration k
k x1 x2 x3 x4
First guess 0 2.0 2.0 2.0 2.0
1
2
3
4
5
6
7
8
9
10

Table 1
Worksheet for the iterative solution.

In some cases, it may be possible to develop improved iteration schemes by using the
latest information that is available. For example, in the above scheme, we could use
the very latest value of x1  when we are calculating x2  , since we already have this
 1  1

quantity. Likewise, we can use both x1  and x2  when we calculate x3  etc, as
 1  1  1

shown below:

©HERIOT-WATT UNIVERSITY December 2017 v2


35

x1
 1
 
 1 3.1  13.92  0.32 x2   0.5 x3   0.1x4
  

x2  1 2.1   5.63   0.20 x 
 1  1
 033x3   0.21x4
 
1

x3 1  1 4.0   15.19   0.23x 1


 1
 0.32 x2 1  3.0 x4  
x4  1 5.2   16.76   0.42 x 
 1  1
 0.22 x2
 1
 0.5 x3
 1
1

The underlined items shows that they are calculated at the latest time available.
Compare the solution using these equations with the original ones. Does the method
converge faster?

Notes on Iterative Schemes


(i) The convergence rate of an iterative scheme may depend on how good the initial
guess is. You can demonstrate this by choosing absurd values for this initial guess
and seeing how long it takes to converge.

(ii) We do not know in advance how many iterations may be required in order to
converge a given iteration scheme. In reservoir simulation, when we solve the
pressure equation, a good guess for the new pressures is usually the values of the
old pressures, at the last time step. If nothing radical has changed in the reservoir,
this is fine. However, if new wells have started up in the model, or if existing wells
have changed rate significantly, the “pressure at the last time step” guess may not
be very good. However, with a robust numerical method, convergence may still
be achieved.

(iii) The equations may not converge – or converge so slowly that it would take too
long to achieve the required accuracy. The iterative scheme described above is
the simplest. In reservoir simulators more sophisticated methods are used.

(iv) As shown above, it helps if the latest information is used.

5.5 Comparison of Iterative and Direct Methods for Solving Numerical


Equations
In the last section we described two general methods - direct and iterative - for solving
the linear equations which arise when we discretise the flow equations of reservoir
simulation. We have not indicated which of these methods is usually “best” for
reservoir simulation problems. This is because, it depends on the problem. In general

©HERIOT-WATT UNIVERSITY December 2017 v2


36

direct methods are best for small problems, but for larger problems, such as the
solution of equations in reservoir simulation, an iterative method is best.

We can summarise our comparison between direct and iterative methods for solving
the sets of linear equations that arise in reservoir simulation as follows:

(i) A direct method for solving a set of linear equations has an algorithm that involves
a fixed number of steps for a given size of problem. Given that the equations are
properly behaved (i.e. the problem has a stable solution), the direct method is
guaranteed to get to the solution in this fixed number of steps. No first guess is
required for a direct method.

(ii) An iterative method on the other hand, starts from a first guess at the solution (x(0))
and then applied a (usually simpler) algorithm to get better and better
approximations to the true solution of the linear equations. If successful, the
method will converge in a certain number of iterations, Niter, which we hope will
be as small as possible. However, we cannot usually tell what this number will be
in advance. Also, in some cases the iterative method may not converge for certain
types of “difficult” problem. We may need to have good first guess to make our
iterative method fast. Also, it often helps to use the latest computed information
that is available (see example above).

(iii) Usually the amount of “work” required for a direct method is smaller for smaller
problems but iterative methods usually win out for larger problems. For an
iterative method, the amount of work per iteration is usually relatively small but
the number of iterations (Niter) required to reach convergence may be large and is
usually unknown in advance.

6. Direct Solution of the Non-Linear Equations of Multi-


Phase Flow

6.1 Introduction to Sets of Non-Linear Equations


We noted in Section 4 above that the equations which we obtain when we discretise
the two-phase flow equations are actually non-linear in nature. However, the strategy
we discussed above (the IMPES method) involved tackling the problem almost as if it
were a linear set of equations since we time-lagged the coefficients to reduce the
problem to a linear set of equations. Then we repeated this process - we applied
repeated iterations - until it (hopefully) converged. Therefore, we took a non-linear
problem and solved it as if it were a series of linear problems.

©HERIOT-WATT UNIVERSITY December 2017 v2


37

In this section, we will introduce the general idea of solving sets of non-linear equations
numerically and indicate how this can be applied to the two-phase flow equations. We
will do this in a simplified manner that shows the basic principles without going into
too much detail. Firstly, compare the difference between solving the following two
sets of two equations, a set of 2 linear equations:

2 x1  x2  10
(59)
3x1  x2  5

and a set of 2 non-linear equations:

x1  2 x2e x1  2
(60)
x2  x1   x1  x2   5
2 2

It is immediately obvious that the second set of non-linear equations is more difficult.
The first set of linear equations can be rearranged easily to show that x1 = 3 and x2 = 4.
However, it is not as straightforward to do the same thing for the non-linear set of
equations. As a first attempt to solve these, we might try to develop an iterative
scheme by rearranging the equations as follows:

 
x1 1  2  2 x2  e  x1

  (61)
2

 1
x2  x1  5
x2 
x1 

Where, as before,  denotes the iteration counter.

The exact solution in this case is x1 = 2.510681 and x2 = 3.144089. Applying the above
iterative scheme with a first guess of x1   x2   1.0 gives the results shown in Table
0 0

2 below (where we have removed some of the iterations).

©HERIOT-WATT UNIVERSITY December 2017 v2


38

Iteration
counter x at iteration v
v x1 x2
First guess 0 1.000000 1.000000
1 2.735759 2.449490
2 2.317673 2.920421
3 2.575338 2.987627
4 2.454885 3.104133
… … …
9 2.513337 3.141814
10 2.508958 3.144173
… … …
25 2.510682 3.144089
26 2.510681 3.144089

Table 2
Iterative method for solving two non-linear equations.

We note that the convergence of these equations is not very fast but, in this case, a
solution is reached. In general, this is not a good approach for solving non-linear
equations.

6.2 Newton’s Method for Solving Sets of Non-Linear Equations


Suppose we have a single non-linear equation such as:
x2e x  0.30 (62)

Which has the solution x = 0.829069 (you can verify this by checking and making sure
the right side is 0.30). This equation can be written as:

f  x   0 where f  x   x 2e x  0.30 (63)

and the solution we require is the value of x for which the function f(x) is zero.

Expand f(x) as a Taylor series as follows:

x 2
f  x   f  x0   xf   x0   f   x0   ... (64)
2

where the dash and double dash represent the first and second derivatives of f,
respectively.

©HERIOT-WATT UNIVERSITY December 2017 v2


39

   1
Thinking in terms of an iterative scheme ( x  x ) as a basis for calculating better
and better solution of f(x) = 0, we can rewrite the Taylor series above as:

  
f  x   f x   x 1  x  f  x     (65)

where we have neglected second order and higher terms. Since we require a solution
at f(x) = 0, we obtain:

  
f x   x 1  x  f  x   0    (66)

which can be rearranged to the following algorithm to estimate our updated guess,
x(+1), as follows:

x  1
x  

 
f x 
f   x  

(67)

This equation is the basis of the Newton-Raphson algorithm for obtaining better and
better estimates of the solution of the equation, f(x) = 0. Note that we need both a
first guess, x(0), and also an expression for f  x    the derivative at iteration . In the

simple example in Equation (62), we can obtain the derivative analytically as follows:

df  x 
 f   x   2 xe x  x 2e x (68)
dx

Therefore the Newton-Raphson algorithm for solving Equation (64) above is:

 x   e  x 
2 

 0.30
x
 1
 x  

(69)
 x 
 e  x 
2
  
  
2x e  x

The point iterative and Newton-Raphson solutions for Equation (62) are shown in Table
3.

©HERIOT-WATT UNIVERSITY December 2017 v2


40

Iteration
Number
Point Newton-
Iteration Raphson

Method Method
x() x()
0 1.000000 1.000000
1 0.903042 0.815485
2 0.860307 0.829011
3 0.842120 0.829069
4 0.834497 0.829069
5 0.831322 0.829069
6 0.830003 0.829069
7 0.829456 0.829069
8 0.829230 0.829069
9 0.829136 0.829069
10 0.829097 0.829069
11 0.829080 0.829069
12 0.829074 0.829069
13 0.829071 0.829069
14 0.829070 0.829069
15 0.829069 0.829069
16 0.829069 0.829069

Table 3
Comparison of the Iterative and Newton-Raphson Methods for a simple problem.

The results in Table 3 show that the Newton-Raphson method converges very rapidly
to the solution, whereas it takes 15 iterations for the point iterative method to
converge at the same level accuracy. The performance can vary quite a bit from
problem to problem but the Newton-Raphson method is generally better if the
   , is not too close to zero. When this derivative gets too small,
derivative term, f  x

it can be seen that the second term in Equation (67) would start to get very large or
“blow up”, as it is sometimes described. We will not go into details about the
convergence properties of the Newton-Raphson but it is meant to be “quadratic”
meaning that the error should decrease quite rapidly.

We now go on to see how the Newton-Raphson method can be applied to sets of non-
linear equations. Returning to the set of two non-linear equations (Equation 60) in the
previous section, we note that another way of writing this set of equations in a general
way is as follows:

©HERIOT-WATT UNIVERSITY December 2017 v2


41

F1  x1 , x2   x1  2 x2e  x1  2  0
(70)
F2  x1 , x2   x2  x1   x1  x2   5  0
2 2

The problem is then to find the values of x1 and x2 that make F1 = F2 = 0. In this section,
we present Newton’s method for the solution of sets of non-linear equations.
Basically, we will simply state the Newton-Raphson algorithm without proof (although
it will resemble the simple form of the Newton-Raphson method above) for sets of
non-linear equations. We will then illustrate what this means for the example of the
set of two non-linear equations above (Equation 60).

Definitions:

As before,

x() and x(+1) = the solution vectors at iteration levels  and ( +1)

New terms are:

N = the number of non-linear equations


(and hence the number of unknowns, x1, x2, … xN)

F(x()) = vector of function values, F1, F2, … FN at x()

F(x) = 0 for the “true” solution of the set of non-linear equations

That is:

 F1  x1 , x2 , x3 ,......, xN  


 
 F2  x1 , x2 , x3 ,......, xN  
 
 F3  x1 , x2 , x3 ,......, xN  
F x 

  (71)
 ..... 
 
 ..... 
 F  x , x , x ,......, x  
 N 1 2 3 N 

For the Newton-Raphson method, we need to differentiate all the Fs with respect to
all the xi, i.e. we require terms of the form: Fi x j , where both i and j go from 1 to

©HERIOT-WATT UNIVERSITY December 2017 v2


42

N. The N x N matrix of differential terms is called a Jacobian matrix, and is defined as


follows:


 
 F1 x    
F1 x 

 
F1 x 

......
   
F1 x 

 x1 x2 x3 xN 


 

 
 F2 x  

 
F2 x 

 
F2 x 

......
 
F2 x 



 x1 x2 x3 xN 
 
 
J x

  F3 x  
 
 
F3 x 

 
F3 x 

 
F3 x 


 ...... 
 x1 x2 x3 xN 
 ... ... ... ...... ... 
 
 ... ... ... ...... ... 

 FN x   
 
FN x 

FN x   

......
 
FN x 
 

 x x2 x3 xN 
 1 
(72)

In the 1D case (only one x value), we had to divide by f   x  , and in the multi-
dimensional case (lots of x values), we need to take the inverse of the Jacobian:

 
1
 J x   .
 

(Recall that the inverse of a matrix, A is the denoted by A1 and by definition,
1
A A  I where I is the identity matrix – all the diagonal terms equal 1 and the other
elements are zero.)

1 0 0 ...... 0
0 1 0 ...... 0 

 0
   
1
 J x   .  J x     I  0 0 1 ......
 (73)
   
... ... ... ...... ...
... ... ... ...... ...
 
0 0 0 ...... 1

©HERIOT-WATT UNIVERSITY December 2017 v2


43

Using the above definitions, the Newton-Raphson algorithm for a set of non-linear
equations, F  x   0 , is given by the following expression:

    
1
x  x    J x
 1
.F x
  

This method could be applied to the Equations shown in (60) and (70) above. In this
case, the Jacobian can be calculated analytically. However, the maths is quite complex,
so is given in the Appendix. (You need not spend time on this, unless you want to try
as a mathematical challenge!) Instead, we simply present the results in the table
below. You can see that the point iterative method does not converge till iteration 26,
but the Newton-Raphson method converges at iteration 7.

Table 4
Comparison of the point iterative method and the Newton-Raphson Method for the
2D example.

©HERIOT-WATT UNIVERSITY December 2017 v2


44

6.3 Newton’s Method Applied to the Non-Linear Equations of Two-Phase


Flow
We now return to the discretised equations of two-phase flow, which we noted at the
time form a set of non-linear equations. The simplified form of these equations is as
follows:

Pressure Equation (Equation 40):

  S 
T
n 1
o
i 1/2
P n 1


 T  Son 1  
i 1/2

  S 
T
n 1
o
i 1/2

 Pi 
n 1

T  Son 1  
i 1/2
Pi n11  0
i 1
x 2
 x 2 x 2
 x 2
 

Saturation Equation (Form B – Equation 44):

  o  So i 1/2 n 1
t    o  Son 1  
n 1

S n 1
oi  S  
n

 
oi
x 2  Pi 1  Pi
n 1
   
  x 2
i 1/2
 Pi
n 1
 P n 1
i 1    0
 
 

The unknowns which we need to determine are Pi n 1 , Sin1 , where i = 1, 2, …, NX. We


can write the equations in a general form as:

FP ,i S  n 1
,P
n 1
0
(74)
FS ,i S n 1
,P
n 1
0
n 1 n 1
The vectors of unknowns, P and S are given by:

©HERIOT-WATT UNIVERSITY December 2017 v2


45

 P1n 1   S1n 1 
 n 1   n 1 
 P2   S2 
 P3n 1   S3n 1 
   
 ...   ... 
 ...   ... 
   
 ...   ... 
P   Pi n11  S   Sin11 
n 1 n 1
(75)
   
 Pi n 1   Sin 1 
 n 1   n 1 
 Pi 1   Si 1 
 ...   ... 
   
 ...   ... 
 ...   ... 
 n 1   n 1 
 PNX   S NX 

However, the functions FP ,i S  n 1


,P
n 1
 and 
FS ,i S
n 1
,P
n 1
 at grid block i only
depend on the quantities Pi n11 , Pi n1 , Pi n11 and Sin11 , Sin1 , Sin11 rather than on all of the
pressure and saturations in the system, since it is only the nearest neighbours which
are coupled together (in the 1D case).

n 1
We may also write all the unknowns in one vector, X , combining all of the
saturations and pressures, as follows:

 S1n 1 
 n 1 
 P1 
 S 2n 1 
 n 1 
 P2 
n 1
 ... 
X   n 1  (76)
 Si 
 P n 1 
 i 
 ... 
 n 1 
 S NX 
 PNX
n 1


©HERIOT-WATT UNIVERSITY December 2017 v2


46

so the equation to be solved is F X  n 1


0.
The saturations and pressures are coupled together to their nearest neighbours
through the discretisation equations (40 and 44 above). The general form of the
solution using the Newton iterations is then to take a starting equation (iteration,  =
n 1 0 
0) X and then update these values using:

   
1
 J X   .F X  
n 1 1 n 1  
X X (77)
 

There are various methods to construct the Jacobian matrix, J  X  . Also, we need to
1
be able to obtain the inverse of the matrix,  J  in order to apply the algorithm. Note
that it is a very sparse matrix because each grid block is only linked to the one on either
side. There are special techniques to invert sparse matrices. We will not go into these
here.

Also note that in MRST, a technique called Automatic Differentiation (AD) is used.
When each variable is calculated, a derivative is also calculated and stored, ready to be
used in the Jacobian.

7. Closing Remarks
In this Topic we have introduced the finite difference approximations for partial
differential equations (PDEs) which describe both single- and two-phase flow through
porous media. The discretised equations lead to systems of either linear or non-linear
equations (depending on whether the coefficients depend on the unknowns), which
are solved numerically. Simple methods for solving the equations have been presented
to give the student a general idea of how reservoir simulation works. However, in most
reservoir simulators the actual methods applied are much more sophisticated because
they are required to deal with models where there may be 100,000s of grid cells (or
more).

8. Appendix
This is a continuation of Section 6.2, giving the full solution of the example with two
non-linear equations. The equations (Equation 62 originally) are:

©HERIOT-WATT UNIVERSITY December 2017 v2


47

x1  2 x2e x1  2
(A1)
x2  x1   x1  x2   5
2 2

For the Newton-Raphson method, they are written as:


F1  x1 , x2   x1  2 x2e  x1  2  0
(A2)
F2  x1 , x2   x2  x1   x1  x2   5  0
2 2

In matrix form:

 F1  x1 , x2    x1  2 x2e 1  2 
x

F  x     (A3)
 F2  x1 , x2    x2  x1   x1  x2   5
2 2

The Jacobian is:

 
 1  2 x  e x1 
  2e   x1 



 
2
 

     x  
J x 2 2  (A4)
 2 x1  x2   x2       
 2 x1 x2 
 
1

Now we need to calculate the inverse of this matrix. This is simple in the case of a 2x2
matrix. If we have a matrix, A , given by

a b
A  (A5)
c d 

It is well-known that the inverse is:

1 1  d b 
A    (A6)
ad  bc  c a 

The inverse on the Jacobian in our example is therefore:

©HERIOT-WATT UNIVERSITY December 2017 v2


48

 
     
1
J x   1

  x   2x    2e   2x  

 
  
 x1 

2 2
 1  2 x2  e  x1      
x2    
x2  x2  
 
1 1 1




1 
 x  2  2 x  x 
1 2   2e  x1 





  2 x  x   x  
1 2 2
2
   1  2 x e  2
  x1 



(A7)

Remember that the equation for the Newton-Raphson Method is:

    
1
x  x    J x
 1
.F x
  
(A8)

Therefore, we can now solve the equations using the following:

 
 1    1 
x 
  x   2x    2e   2x  
x 
  
 x1 

2 2
 1  2 x2  e  x1      
x2    
x2  x2  
 
1 1 1




1 
 x  2  2 x  x 
1 2   2e   x1 


 
 
x1   2 x2  e  x1  2

       2
 

 
  2 x  x    x  
1 2 2    1  2 x e 
2

2
 x1 
 
  x2 x1
 
   
 x1 x2     2
 5

The results are shown are in Table 4 in Section 6.2.

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 6 – Modelling and Upscaling


Contents
1. Introduction ............................................................................................................ 4
1.1 Nomenclature for Grid Structure ......................................................................... 4
1.2 Length Scales ........................................................................................................ 5
1.3 Brief Overview of Geological Modelling Approaches ........................................... 6
1.3.1 Assigning Grid Block Properties ..................................................................... 7
1.3.2 Software for Generating Stochastic Models ............................................... 11
1.3.3 Two Reasons for Upscaling .......................................................................... 12
2. Averaging Methods .............................................................................................. 14
2.1 REV – Representative Elementary Volume ........................................................ 14
2.2 Averaging Porosity and Water Saturation .......................................................... 15
2.3 Averaging Permeability ...................................................................................... 16
2.3.1 Flow Parallel to Uniform Layers................................................................... 18
2.3.2 Flow Perpendicular to Uniform Layers ........................................................ 19
2.3.3 Flow in a Correlated Random Permeability Field ........................................ 21
2.3.4 Additional Averaging Methods .................................................................... 23
2.3.5 Summary of Permeability Averaging ........................................................... 23
3. Numerical Upscaling Methods ............................................................................. 25
3.1 Flow Equations for an Incompressible System ................................................... 25
3.2 Numerical Upscaling with No-Flow Boundary Conditions ................................. 26
3.3 Accuracy of Upscaling for Single Phase Flow ..................................................... 29
4. Two-Phase Flow .................................................................................................... 30
4.1 Introduction ........................................................................................................ 30
4.2 Applying Single-Phase Upscaling to Two-Phase Problems ................................. 35
4.3 Improving Method of Single-Phase Upscaling.................................................... 38
5. Two-Phase Upscaling Methods ............................................................................ 40
5.1 Introduction to Two-Phase Upscaling ................................................................ 40
5.2 Steady-State Methods ........................................................................................ 41

©HERIOT-WATT UNIVERSITY December 2017 v2


2

5.2.1 Capillary Equilibrium.................................................................................... 41


5.3 Two-Phase Dynamic Upscaling ........................................................................... 42
5.3.1 Outline of Kyte and Berry Method .............................................................. 42
5.3.2 Discussion on Numerical Dispersion ........................................................... 43
5.3.3 Disadvantages of Dynamic Two-Phase Upscaling ....................................... 45
6. Scale Issues ........................................................................................................... 46
6.1 Effect of Small-Scale Structures.......................................................................... 47
6.2 The Geopseudo Method ..................................................................................... 49
6.2.1 When to Use the Geopseudo Method ........................................................ 50
7. Streamline Simulation .......................................................................................... 51
8. Concluding Remarks ............................................................................................. 53
9. References ............................................................................................................ 54
10. Appendix – The Kyte and Berry (1975) Method ............................................... 55

©HERIOT-WATT UNIVERSITY December 2017 v2


3

Learning Objectives
After studying this topic, you should be able to:

 Outline the procedure for geological modelling.

 Calculate statistical properties for poro-perm data.

 Describe why we need to upscale data for reservoir simulation.

 Explain the concept of REV (representative elementary volume).

 Calculate the average porosity and water saturation.

 Calculate the average permeability of simple models


o Prove that the average permeability for flow along parallel layers is the
arithmetic average.
o Prove that the average permeability for flow across parallel layers is the
harmonic average.
o Use the geometric average to calculate the average permeability in a random
distribution.

 Outline the procedure for calculating the effective permeability numerically.


o List common types of boundary conditions.

 Discuss cases where upscaling errors are likely to be large.

 Demonstrate the effect of heterogeneity on two-phase flow.

 Outline methods for upscaling two-phase flow.

 Describe the effects of small-scale structure on two-phase flow.


o Describe the Geopseudo Method and discuss when to use it.

 Outline the procedure for streamline simulation.


o Discuss the advantages and disadvantages of streamline simulation.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

1. Introduction
In this Topic, we present two related procedures which are required for reservoir
simulation, namely modelling and upscaling. For numerical simulations, some sort of
model (grid structure) is always required. (See Topic 3, Section 2.) This may be a very
simple grid. For example, you may wish to test the miscibility of a gas and oil for a gas
flood. In this case, you might just use a simple Cartesian (perhaps 1D) grid, with
negligible geological input. On the other hand, it is usual nowadays to build a 3D corner
point grid for a reservoir to predict future recovery and evaluate various sensitivities.
In both cases, we need to decide how fine or coarse we require the model to be. Also,
in the case of a geological model, we need to decide how to populate the cells with
attribute values (porosity, permeability, etc). Figure 1 shows a diagram of the different
types of data required for reservoir modelling and simulation. (This is the same as
Figure 7 in Topic 1.)

Figure 1
Data required for reservoir modelling and simulation. The shaded blocks represent
data which will be mentioned in this Topic.

1.1 Nomenclature for Grid Structure


Before discussing modelling further, we specify the nomenclature which will be used
to describe a grid. Assume for the moment, that we have a Cartesian grid, and that
the total size of the model is Lx  Ly  Lz (measured in ft or m). This is split up into
grid cells of size dx  dy  dz (measured in ft or m), so that there are nx  ny  nz cells,
where Lx  nx  dx , etc. Note that, even if you are only using a 1D grid, you still need
to specify the size of the cells in each dimension (i.e. dx, dy and dz). See Figure 2. In a

©HERIOT-WATT UNIVERSITY December 2017 v2


5

corner point grid, we can still use the same nomenclature, but the grid cells will have
uneven shapes and sizes.

We might want to make a very fine grid which can resolve a lot of geological detail, but
this might take too long for flow simulation. Therefore, as mentioned in Topic 1, we
need to compromise. Also, if there is little data, it will not be worthwhile making a
fine-scale model. In a coarse model, we will need to somehow average properties such
as porosity and permeability. In other words, we need to upscale the data.
Additionally, sometimes geologists construct too fine a model for flow simulation, so
engineers have to perform another stage of upscaling.

dy
dz dx
Lz
Ly
Lx
Figure 2
Nomenclature for a grid. In this case, nx = 7, ny = 5 and nz = 3.

1.2 Length Scales


As discussed in Topic 1, reservoirs are usually very complex with heterogeneities at
many different scales. Figure 3 shows typical length scales which arise in a sandstone
reservoir. The x-axis shows horizontal length-scale and the y-axis shows vertical length
scale. Note that the axes are in a logarithmic scale, and that the scales are much larger
in the horizontal direction. At the mm-cm scale there are laminae. These group
together to form beds over scales of several meters to 100m in the horizontal and
about a meter in the vertical. Then beds are grouped into sequences and para-
sequences. The length-scales for different types of measurement are also shown on
the graph, as well as the size of typical full-field geological models and simulation
models.

In Figure 3 you can see that the effects of lamination cannot be resolved by core and
log data, only by fine-scale probe permeameter measurements, and this data may not
be available. The fine-scale lamination may not be important, but sometimes can have

©HERIOT-WATT UNIVERSITY December 2017 v2


6

significant effects on two-phase flow. If so this should be taken into account. (See
Section 6.2.)

1.3 Brief Overview of Geological Modelling Approaches


In this simple review of modelling, we are going to start by assuming that the reservoir
structure has been given (from seismic data, or we have decided that a Cartesian grid
is adequate). The steps for creating a model are:
a) Decide on the grid resolution
 Remember that models should be “fit-for-purpose”. You should not make a
very fine model if it is not required, or if you do not have sufficient data.
 For the moment let’s assume that we are going to model at the scale of the
“geological model” in Figure 3, i.e. with grid cells of about 100 m in the
horizontal and 1 m in the vertical.

b) Acquire data for the model – porosity, permeability, etc, from core plugs and well
logs
 We only have data measurements at the wells (apart from some attributes
from seismic data).
 If core plug measurements have been made, we will have porosity and
permeability measurements at the scale of a core plug (a few cm). Such
measurements can give us correlations between log(permeability) and
porosity as shown in Topic 7 of Course G19RA.
 There is usually more information from well logs, although it will be at a
coarser resolution. If porosities have been obtained from the neutron-density
log, these can be used to calculate permeabilities. (However, note that the
poro-perm relationship at the plug-scale may not be the same as the poro-
perm relationship at the well-log scale, depending on the heterogeneity.)

c) Populate grid at the wells with poro-perm values.


 Since the grid block size is greater than the scale at which the measurements
were made, some form of upscaling of the values is required, usually
averaging. However, upscaling of permeability is a complex topic, which is why
most of this Topic is devoted to permeability upscaling.

d) Decide how to populate the grid with values, away from the wells.
 This may be done deterministically or stochastically. We focus on this below.
 Often the rock type is assigned first, and then the properties are generated for
each type of rock.

©HERIOT-WATT UNIVERSITY December 2017 v2


7

From Pickup and Hern (2002), adapted by Barkve (2004)


100

10 Seismic data Para-


sequences
Vertical thickness (m)

Log
1 Flow model

0.1 Core Geological model

Beds
0.01 Probe

Laminae
0.001
0.001 0.01 0.1 1.0 1 10 100 1000 10000
Horizontal length (m)

Figure 3
Length scales of sedimentary structures, data and models.

1.3.1 Assigning Grid Block Properties


There are two ways in which properties may be assigned to a model:
a) Deterministically
b) Stochastically (or probabilistically)

In a deterministic model, we specify the porosity and permeability precisely. Examples


include:
 a homogeneous model in which the porosity and permeability are the same in
every cell (e.g.  = 0.2, k = 100 mD.)
 a layered model where the porosity and permeability are constant within each
layer, but the values in each layer are different. This type of model could be used
if we can correlate layers between wells, and if the permeabilities in the layers are
relatively homogeneous (Figure 4a).
 a model which represents some form of sedimentary structure (e.g. cross-bedding)
where different laminae are assigned high or low poro-perm values. This kind of
model might be constructed if small-scale structure is likely to have a significant
effect on waterflood performance (Figure 4b).

©HERIOT-WATT UNIVERSITY December 2017 v2


8

Figure 4
Examples of deterministic models: a) A layered model. The size of a grid cell could be
~ 100m x 1 m. b) A model of laminae. In this case, the cell size may be only a few
mm.

One the other hand, we know that rock properties are very heterogeneous, and it is
difficult to predict the properties between wells. For this reason, many people
generate stochastic models. This means that there is a random variation in the
properties. The branch of geological modelling which deals with stochastic methods is
called geostatistics.

We shall not go into geostatistics in detail. However, the student should become
familiar with the following terms (if they are not familiar with them already).

Assume we have uniformly spaced measurements xi, where i = 1, n.


n

x i
The mean, or average value is:   i 1
. (1)
n

The standard deviation, which is a measure of the spread of the data, is:
 xi   
2

  . (2)
n

The variance is the square of the standard deviation, so:


x  
2

  i
2
. (3)
n

Many properties have an approximately normal (or Gaussian) distribution of values,


i.e. the probability distribution, f(x), is given by:

©HERIOT-WATT UNIVERSITY December 2017 v2


9

1   x   2 
f  x  exp   . (4)
2 2  2 2 

Figure 5 shows examples of normal distributions. In a reservoir, porosity may be


normally distributed. However, permeability is more likely to be log-normally
distributed – i.e. the x-axis is log(k), not k. Figure 5 c) shows the permeability
distribution when the permeability is log-normally distributed. Note that the mean of
ln(k) = 4.605, which is equivalent to 100 (mD).

Figure 5
Examples of normal distributions. a) a normal distribution of porosity; b) A normal
distribution of log(permeability); c) the equivalent permeability distribution.

We could generate stochastic values (e.g. porosities) for the grid blocks in a model, if
we assumed that the values were normally distributed, by specifying the mean and
standard deviation. However, using this approach, the value for each grid cell would
be independent of surrounding grid blocks, and this is unrealistic. In a reservoir
adjacent samples of rock are likely to be more similar than samples which were
measured a distance apart. This is referred to as spatial correlation. One way of
measuring the spatial variation of measurements is to calculate a variogram. In fact,
usually a semi-variogram is calculated. This is done by comparing pairs of values a
certain distance apart (called the lag). A variogram can be calculated if the
measurements are irregularly spaced. However, it is easier to demonstrate if they are

©HERIOT-WATT UNIVERSITY December 2017 v2


10

evenly spaced – e.g. the porosity in core plugs, measured every foot, over an interval
of several hundred feet.

If the lag is 1 ft, then compare pairs 1 and 2, 2 and 3, 3, and 4, and so on.
If the lag is 2 ft, then compare pairs 1 and 3, 2 and 4, 3 and 5, etc. See Figure 6.

1 2 3 4 5 6 7 8

Figure 6
Pairs of measurements used to calculate a semi-variogram. The blue dots represent
data points, and the curved lines are links between pairs. This figure only shows the
links for lags of 1 and 2. To make a whole semi-variogram, more lags are required.

The formula for the semi-variogram is:

np

 x  xi  j  ,
1
  h 
2
i
(5)
2n p i 1

where h is the lag (in units of ft of m)


j is the point a distance h from i (j = 1, 2, 3 …),
np is the number of pairs (np = the total number of pairs = n - j).

An example of a semi-variogram is shown in Figure 7. In general the values start small


at low lags, and increase until a plateau is reached. The plateau is referred to as the
sill, and is equal to the standard deviation squared (i.e. the variance). The lag at which
the sill is reached is called the range. The nugget is the value of the semi-variogram
extrapolated to zero, and represents heterogeneity at a smaller scale than the
measurements.

The semi-variogram is usually different in each direction, because the range is


different: it is much longer in the horizontal direction. Also, the poro-perm variation
in the vertical direction may be cyclic, giving rise to a “hole” (dip) in the semi-
variogram. This is because at a certain scale, the data are more similar (e.g. at the scale
of beds).

©HERIOT-WATT UNIVERSITY December 2017 v2


11

semivariogram sill

nugget range

Lag (m)
Figure 7
An example of a semi-variogram. The red dots are data and the black curve is a fit to
a model (spherical variogram).

1.3.2 Software for Generating Stochastic Models


There are several methods for generating stochastic models, or correlated random
fields, as they are often called. The Petrel software package can be used to generate
such models. It is a complex package this allows the user to:
 Input seismic to model the reservoir structure
 Input well log and core data
 Calculate semi-variograms for well log and core data
 Grid the model
 Assign properties to the grid – e.g. rock type, net:gross, porosity, permeability, etc,
using stochastic methods, but constraining the values at the wells.

One method which Petrel uses for generating correlated random properties is known
as Sequential Gaussian Simulation (SGS).

Note that, since porosity and log(permeability) are usually correlated, their values
should not be generated independently. Instead, the porosity is usually generated as
a correlated random model, and then the permeability in each grid cell is calculated
from the porosity value in that cell. This may be done assuming that there is perfect
correlation between porosity and log(k). For example an empirical relationship may
be used. The example in Figure 5 has the following relationship:

ln  k   25  0.395 .

Alternatively, a random variation may be superimposed on this relationship to allow


for the fact that there is not perfect correlation between them. This may also be done
in Petrel.

©HERIOT-WATT UNIVERSITY December 2017 v2


12

In MRST, a simpler method is used to generate correlated random models. A Gaussian


(normal distribution is generated and is “convolved” (spread out) with a Gaussian-
shaped function to correlate the values (Lie, 2015, P64-65). Figure 8 shows an example
from the MRST book (Lie, 2016). It shows the porosity distribution for a 2D model of
dimensions 50 x 20.

Figure 8
Example of a correlated random model generated in MRST. a) porosity distribution;
b) histogram of porosity. (From Lie, 2016.)

1.3.3 Two Reasons for Upscaling


Upscaling is used at different stages of the modelling and simulation process:

a) Upscaling of data to go into a model


As mentioned above, grid cells may be larger than the scale at which the
measurements were made. Suppose you are using log data (e.g. porosities), and that
you have a value every 0.1m. If you are constructing a geological model with a grid cell
of thickness of 0.5 m, you will have to average 5 values together to give a value for the

©HERIOT-WATT UNIVERSITY December 2017 v2


13

grid cell at that location. This is a valid approach if you assume that the measurements
are representative of that part of the reservoir. On the other hand, this part of the
reservoir may be part of a complex structure, so averaging may not give a good answer,
in which case, you should create a very fine-scale geological model. (See Section 6.2.)
Figure 9 shows an example of upscaling of well data. The arithmetic average was used
(assuming a layered model).

b) Upscaling of fine models


As mentioned above, geologists often construct a detailed reservoir model which has
too many grid cells for flow simulation. In this case, the geological model must be
upscaled to the flow model. In most of the examples below we consider upscaling
from a fine grid to a coarser grid. See Figure 9Error! Reference source not found.b).
The structure in a fine-scale model may be complex, so numerical upscaling methods
are often used for this stage of upscaling. However, in some cases averaging may be
used. We discuss these cases in the next section.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Full-field model
Geological model

300,000 cells

20 million cells

Figure 9
Upscaling examples. Top: Example of upscaling well data using the arithmetic
average. Bottom: Example of upscaling a geological model for full-field simulation.

2. Averaging Methods

2.1 REV – Representative Elementary Volume


The diagram in Figure 3 shows a range of length-scales for features such as laminae or
beds. However, within a single rock type, there tends to be specific length scales, and
when modelling a reservoir, the geologist and engineer should be aware of these
scales.

Imagine we are able to make measurements at a range of length scales – e.g.


measuring porosity by sampling different volumes of rock starting at a length-scale of
about 10 microns (m). See Figure 10. The sample may contain either a pore ( = 1),
a grain ( = 0) or part grain, part pore. As we take samples at larger scales the value of
porosity will fluctuate depending on the proportion of grains and pores that are in the
sample. However, the fluctuations will gradually decrease as the sampling volume
increases, until each sample contains a representative number of grains and pores.
When the fluctuation dies down, we have a representative elementary volume (REV),
and the rock may be considered as homogeneous at this length-scale. Of course, as
the sampling size gets even larger, the fluctuations may begin again as different rock
types are included (e.g. sampling part sandstone, part shale). When upscaling, we
want to identify the REVs of the particular rock formations, and upscale to these
length-scales. For example: lamina-scale, bed-scale, bed-set scale. Usually geologists

©HERIOT-WATT UNIVERSITY December 2017 v2


15

do not start modelling at the lamina scale. However lamination may give rise to
significant effects. See Section 6.2 on the Geopseudo Method.

Figure 10
Schematic diagram of how a property may vary with the volume of the sample. From
Nordahl and Ringrose (2008), re-drawn from Bear (1972).

The concept of REVs is useful when considering length-scales for modelling, especially
for single-phase flow. In the following Section, we shall assume that we are upscaling
to an REV, so that upscaling is reasonably accurate. We shall return to the topic of
accuracy in Section 3.3.

For the moment, let us assume that a fine-scale grid has been generated, and we are
upscaling to a coarser grid. Figure 11 shows a fine grid split into coarse cells (outlined
in bold). The cells within each coarse block are upscaled in turn.

fine grid coarse grid

one coarse cell

Figure 11
Upscaling from a fine grid to a coarse one.

2.2 Averaging Porosity and Water Saturation


Quantities such as porosity, net-to-gross (the proportion of a grid cell which is
permeable rock) and water saturation are additive, and so are easy to upscale. These

©HERIOT-WATT UNIVERSITY December 2017 v2


16

will be dealt with first. Figure 12 represents part of a fine-scale grid whose cells have
volume Vi, porosity, i, and water saturation, Swi, where i = 1, 2, … n, the number of
cells.

Vi, i, Swi


Figure 12
Example of some fine-scale cells which are to be upscaled to a single coarse cell.

Total Net Volume V NTG i i


Average net-to-gross, NTG   i 1
n
. (6)
TotalGross Volume
V
i 1
i

Total PoreVolume V NTG i i i


Average porosity,    i 1
n
. (7)
TotalVolume
V NTG
i 1
i i

Total Volume of Water V NTG  S i i i wi


Ave. water saturation, S wi   i 1
n
. (8)
TotalPore Volume
V NTG 
i 1
i i i

Note that these formulae are absolutely correct: there are no errors in upscaling
NTG,  or Sw (assuming that the fine-grid values are “correct”). In the following
sections, we shall assume that NTG = 1.

2.3 Averaging Permeability


On the whole, it is not correct to simply average the permeability: we should calculate
an effective permeability, which is defined as:
The effective permeability of a heterogeneous model is the permeability of a single
homogeneous cell which gives rise to the same flow as the fine-scale model when the
same pressure gradient is applied.

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Note that here “effective” means taking account of heterogeneity. In reservoir


engineering, “effective permeability” is sometimes used to specify the permeability of
one phase in the presence of another.

P P

Q Q
keff

kij
Figure 13
Example of effective permeability.

Also note that in upscaling, we are assuming that Darcy’s Law holds both at the fine-
and coarse-scales. For example, flow between grid cells may be represented by
Equations (9) and (10), where lower-case letters are used for the fine scale and upper-
case for the coarse scale. (Compare with equations for flow between grid cells in Topic
3.)

ki 1/2, j a  pi 1  pi 
Fine scale: qi 1/2, j   (9)
 x

keff , I 1/2, J A  PI 1, J  PI , J 


Coarse scale: QI 1/2, J   (10)
 X

True effective permeability (calculated at the scale of an REV) is an intrinsic property


of the model and ought to be independent of the applied boundary conditions.
However, in practice, the effective permeability often does depend on the boundary
conditions, and on the method used for calculation. Upscaling must always be carried
out with care in order to obtain “sensible” results.

In this section, we assume that there is only one phase present – water or oil, and that
we have steady-state linear flow. (We are not considering radial flow, or transient
flow.) We show how simple averaging may sometimes be used to estimate upscaled
permeability. Then in the next section we move on to methods which involve
numerical simulation.

©HERIOT-WATT UNIVERSITY December 2017 v2


18

2.3.1 Flow Parallel to Uniform Layers


If we have infinite, uniform parallel layers, with flow along the layers, it can be shown
that the arithmetic average gives the correct value for the effective permeability. In
fact, we use a thickness-weighted average of the permeability.

Figure 14
Flow along uniform layers.

Proof:
Assume we have n layers, labelled i = 1, 2, .., n, and that the permeability and thickness
of the ith layer are ki and ti, respectively. Also assume that the pressure is P1 at the
left side of a coarse block and P2 at the right side. Then the flow, Qi, will depend on
the permeability and thickness of layer i. We may write Darcy’s Law for each layer:

ki ti y  P 2  P1
Qi   (11)
 x

where y is the (constant) width of the model.

The total flow through all the layers is:

n n
 k t y  P 2  P1  y  P 2  P1 n
Q   Qi    i i   ki ti (12)
i 1 i 1   x   x i 1

But, the total flow through the model is also

keff T y  P 2  P1
Q , (13)
 x

©HERIOT-WATT UNIVERSITY December 2017 v2


19

n
where T   ti .
i 1

y  P 2  P1 n k T y  P 2  P1
Therefore: Q  
 x
i 1
kiti   eff
 x
, (14)

Cancelling constant terms, we get:

n n

k t
i 1
i i  keff T  keff  ti
i 1
(15)

k t i i
Therefore, the effective permeability is keff  i 1
n
. (16)
t
i 1
i

In other words, the effective permeability is the thickness-weighted average.


Frequently, if we have an approximately layered model, and the layers are more or less
horizontal, we use the arithmetic average as an approximation for the effective
permeability.

2.3.2 Flow Perpendicular to Uniform Layers


If we take the same model, but apply a pressure gradient across the layers, so that the
flow is perpendicular to the layers, the harmonic average gives the correct value for
the effective permeability. Again, we take a thickness-weighted average of the
permeability.

Figure 15
Flow across uniform layers.

©HERIOT-WATT UNIVERSITY December 2017 v2


20

Proof:
Use the same notation as before. This time note that the total flow Q is the same
across all the layers, but the pressure drop across each layer is different.

ki xy Pi
Darcy’s Law for flow across layer i is: Q   (17)
 ti

Qti
Therefore the pressure drop is: Pi   (18)
ki xy

The total pressure drop is therefore:

n n
Qti Q n
 P1  P2    Pi     ti k i (19)
i 1 i 1 ki xy xy i 1

But this can also be expressed as:


n
Q  ti
n
QT
 P1  P 2    Pi    i 1
(20)
i 1 keff xy keff xy

Therefore:

n
Q  ti
Q n
i 1

keff xy
  ti ki
xy i 1
(21)

Cancelling Q, , x and y, and rearranging gives:

t i
keff  n
i 1
(22)
 ti k i
i 1

which is a thickness-weighted harmonic average.

Example
Suppose that the thicknesses and permeabilities of the layers have the values given
below (Figure 16). What is the effective permeability in the horizontal and vertical
directions?

©HERIOT-WATT UNIVERSITY December 2017 v2


21

Figure 16
Permeability and thickness values for the numerical example.

The effective permeability for flow along the layers =

500  30  100  45  20 10  600  35 40700


  339.2 mD .
30  45  10  35 120

The effective permeability for flow across the layers =


30  45  10  35 120
  112.3mD .
30 500  45 100  10 20  35 600 1.06833

The harmonic average is always smaller than the arithmetic average. This can be
understood by considering the effect on a single low permeability layer. If flow is
parallel to the layers, a low permeability layer will have little effect. However, if the
flow is across the layers, a low permeability will have a significant affect. Try changing
the value of the low permeability layer in the example above from 20 mD to 1 mD.
How much effect does this have on the arithmetic and harmonic averages?

2.3.3 Flow in a Correlated Random Permeability Field


The proofs for the effective permeability of layered models is simple. However, the
derivation of a formula for the calculation of the effective permeability of a correlated
random model is more complex, so we shall not show the derivation here. In this case
the effective permeability depends on the geometric average.

 n   n 
  ln  k i     log10  ki  
k geom  exp  i 1   10 ^  i 1 . (23)
 n   n 
   
   

©HERIOT-WATT UNIVERSITY December 2017 v2


22

Figure 17 below shows an example of a model of correlated random permeabilities.


The grid is 30 by 10 cells and the colours indicate log10(permeability). Assume that the
cells are numbered i = 1, 2, …, n (where n = 30 x 10 = 300).

Figure 17
Example of a correlated random permeability model.

In fact, it can be proved that the effective permeability of a correlated random model
depends on the dimension of the model (1D, 2D or 3D). The results given below have
been derived theoretically for log-normal distributions, with a standard deviation of
Y, where Y = ln(k).

keff  k g 1   Y2 2  in1D
keff  k g in2D (24)

keff  k g 1   Y2 6  in3D

These formulae are approximate, and assume that Y is small (< 0.5). The 1D result is
an approximation of the harmonic average. Note that the results do not depend on
the correlation length of the field, provided it is much smaller than the system size.
Also note that ka > kg > kh, and the effective permeability always lies between the two
extremes: ka and kh.

Example
Imagine that the permeabilities in the example shown in Figure 16 were jumbled at
random, but the numbers of cells of each permeability were in the same ratio (i.e.
30:45:10:35). What would be the effective permeability of this model?

©HERIOT-WATT UNIVERSITY December 2017 v2


23

 30 log(500)  45log(100)  10 log(20)  35log(600) 


keff  k geom  10 ^  
 30  45  10  35 
 281.21 
 10 ^    10
2.3435
 220.5 mD.
 120 

2.3.4 Additional Averaging Methods


Since averaging is very quick (compared with numerical simulation), many engineers
use this technique in more complex models. Sometimes engineers increase the
accuracy by using power averaging. The power average is defined as:

1/
 n 
  ki 
k p   i 1  , (25)
 n 
 

where  is the power. The value of the power depends on the type of model, and
must be calibrated against numerical simulation (Section 3.2).

Also, sometimes, engineers use a combination of the arithmetic and harmonic


averages, e.g. they take the arithmetic average of the permeabilities in each column
and then calculate the harmonic average of the columns.

2.3.5 Summary of Permeability Averaging


To summarise, there are two types of simple model in which we can calculate the
effective permeability by averaging:

 Parallel layers
 Correlated random fields

Since averaging is very quick, it is frequently used as an approximation for the effective
permeability in more complex models.

Example 1
Figure 15.1 in Lie (2016) presents a good illustration of upscaling of porosity.
(Reproduced in Figure 18 below.) Note how the histogram of the porosities changes
with upscaling. You can try this example in the Tutorial, and also compare the different
types of average for permeability.

©HERIOT-WATT UNIVERSITY December 2017 v2


24

Figure 18
Example of upscaling porosity. (From Lie, 2016.)

Example 2
In the Tutorial, you can also try upscaling a 2D log-normal permeability field, and
comparing the fine- and coarse-scale permeability histograms.

Figure 19
Example of upscaling a 2D random model using the geometric average.

©HERIOT-WATT UNIVERSITY December 2017 v2


25

3. Numerical Upscaling Methods


In general, the permeability distribution will not be simple enough for us to be able to
calculate the effective permeability analytically (i.e. by averaging), and we will have to
perform a numerical simulation. This is also known as “flow-based upscaling”. We can
use a finite difference method to perform flow simulation.

3.1 Flow Equations for an Incompressible System


In Topic 4 we derived the flow equations using mass conservation (continuity equation)
and Darcy’s Law. Usually when upscaling we make the assumption that we have
incompressible rock and fluids (otherwise the effective permeability would not be a
constant property), and also that the system is in a steady state (nothing changes with
time). This means that a volume balance may be used instead of a mass balance.
Consider part of a 2D model, as shown in Figure 20. The continuity equation tells us
that there is no net accumulation or loss of fluid within a grid block:

qxin  qzin  qxout  qzout (26)

Figure 20
Part of a 2D grid for demonstrating the equations for incompressible flow.

Darcy’s law is used to express the flows in terms of the pressures and permeabilities.
For example, if the grid blocks in Figure 20 are of length x, width y (into the paper)
and height z, then:

k x ,i 1/2, j yz  Pi , j  Pi 1, j 


qxin   (27)
 x

where kx,i-1/2,j is the harmonic average of the permeabilities in the x-direction in blocks
(i-1,j) and (i,j). (You now should know now why the harmonic average is used here.)
The other flows are calculated in a similar manner.

©HERIOT-WATT UNIVERSITY December 2017 v2


26

It is useful to use the transmissibilities, Tx = kxyz/x and Tz = kzxy/z. We can


therefore derive the pressure equation:

T x ,i 1/2, j  Tx ,i 1/2, j  Tx ,i , j 1/2  Tx ,i , j 1/2  Pi , j


Tx ,i 1/2, j Pi 1, j  Tx ,i 1/2, j Pi 1, j
(28)
Tx ,i , j 1/2 Pi , j 1  Tx ,i , j 1/2 Pi , j 1
0

This is left as an exercise to the student. (Note that  cancels in the equation.) See
Topic 5, if you have problems.

An equation is set up for each Pij, i = 1, 2, .. nx and j = 1, 2, .. nz. (Usually we number


the cells as m = 1, 2 …, nx x ny, as in Topic 5, Section 3.3.2.) The transmissibilities are
known, and using the appropriate boundary conditions, we can solve this set of linear
equations to obtain the pressure in each grid block. The effective permeability is then
calculated from the total flow and the total pressure drop, as described below.

Note that the boundary conditions are applied to each coarse grid cell in turn (referring
to Figure 11), and they may not be a good approximation to the pressures which would
arise in a fine-grid simulation of a whole field. This leads to errors in the results.
Upscaling errors are discussed in Section 3.3.

3.2 Numerical Upscaling with No-Flow Boundary Conditions


As in Topic 5, we need to specify boundary conditions to be able to solve the set of
equations. These may be pressures or flows at the edge of the model. It is important
to be aware of what boundary conditions are used, because the result of a flow
simulation depends on the boundary conditions. Obviously there are many ways in
which we could specify the boundary conditions. Here we describe one commonly
used type of boundary conditions, which may be described as “fixed pressure” and/or
“no flow” boundary conditions (Figure 21).

©HERIOT-WATT UNIVERSITY December 2017 v2


27

Figure 21
Fixed pressure, no-flow boundary conditions.

The pressure is fixed on two sides of the model, and no flow is allowed through the
others sides. This type of boundary condition is suitable for models where there is little
cross-flow: for example, models with approximately horizontal layers, or a random
distribution (assuming that the coarse cell is a REV). These are the most commonly
applied boundary conditions. Figure 22 illustrates how an effective permeability may
be calculated in the x-direction.

1. Solve the steady-state equation to give the pressures, Pij, in each grid block
(Equation 28).
2. Calculate the inter-block flows in the x-direction using Darcy’s Law. (See Equation
27.)
3. Calculate the total flow, Q, by adding the individual flows between two y-z planes.
(Any two planes will do, because the total flow is constant.)
4. Calculate the effective permeability for flow in the x-direction, using the equation:

keff , x A  P1  P 2 
Q (29)
L
Repeat the calculation for flow in the y- and z-directions, to obtain keff,y and keff,z.

©HERIOT-WATT UNIVERSITY December 2017 v2


28

Figure 22
The calculation of effective permeability using no-flow boundary conditions.

As mentioned above there are many other types of boundary conditions – e.g. fixing
the pressure at each end and letting the pressure vary gradually along the sides. The
resultant effective permeability usually depends on the boundary conditions used, but
the actual boundary conditions in a reservoir might not be known. Therefore,
sometimes engineers apply the boundary conditions to a region which is larger than
the coarse block. This is known as a “flow jacket”, or “skin” (Figure 23).

Figure 23
Example of a flow jacket round a coarse cell. In this case the flow jacket is 4 cells
thick.

Examples of numerical upscaling are given in the section below, which discusses the
accuracy of upscaling.

©HERIOT-WATT UNIVERSITY December 2017 v2


29

3.3 Accuracy of Upscaling for Single Phase Flow


If we have single-phase flow (just water or just oil) and we have upscaled to the size of
an REV, then the upscaling should be accurate. The example below (Figure 24a) shows
a 2D correlated log-normal permeability distribution which has been upscaled by
different factors. The fine-scale model is 120 x 120 grid cells, each of size 1m x 1m x
1m. This model can be upscaled by a range of different factors:
3x3, 4x4, 5x5, 8x8, 10x10, 12x12, 15x15, 16x16, 20x20, 24x24, 30x30, 40x40. In
upscaling, the aim (as mentioned in Section 2.3) is that there should be the same flow
through the model at the coarse-scale as there was at that fine-scale, when the same
pressure gradient is applied. Therefore, in this model, the total flow through the model
was calculated for the fine-scale grid and for the various course-scale grids. If upscaling
works perfectly, the flows should be the same in the fine- and coarse-scale models.
This means that the ratio of the flows should be 1. Figure 24b) shows that the flows
are similar, so there is little error in upscaling this model. However, there is a dip in
the plot around a scale-up factor of 10, which is approximately the range of the semi-
variogram. This example can be tried in the Tutorial.

The results here are typical of this type of model, where the accuracy is high for a very
small scale-up factor (little change in the permeability distribution), and also the
accuracy is high for a large scale-up factor (which includes an REV), but the accuracy is
less at intermediate scale-up factors. This result only applies for single-phase
systems.

Figure 24
Example of checking numerical upscaling by comparing the ratio of coarse-scale to
fine-scale flows: a) permeability distribution and b) plot of flow ratio.

©HERIOT-WATT UNIVERSITY December 2017 v2


30

In many cases upscaling is much more inaccurate, particularly in models with very high
permeability contrasts (i.e. many reservoir models!), and there may not be an obvious
REV to upscale to. Examples of models where upscaling is tricky are:
 Low permeability shales in a high permeability sandstone
 Low permeability faults in a high permeability sandstone
 High permeability channels in a low net/gross reservoir
 High permeability fractures in a low permeability reservoir.

All these cases are difficult to model. As an example, we consider a sand/shale model,
where the shale has zero permeability. Figure 25 shows part of a fine-scale model,
which has to be upscaled to 3 coarse blocks, as shown. Since there is a shale lying
across each coarse block, each coarse block will have zero permeability in the z-
direction (vertical). However, fluid can flow through the model vertically, as shown.
This error arises because the coarse block size is similar to the characteristic length of
the shales. Upscaling would be more accurate, if the coarse block size was much larger,
or much smaller than the shales. Alternatively, using a “skin” or “flow jacket” will
increase the accuracy of upscaling (Figure 23).

Figure 25
Sand/shale model.

4. Two-Phase Flow

4.1 Introduction
Often we need to simulate two-phase systems, e.g. a water flood or a gas flood of an
oil reservoir, or an oil reservoir with a gas cap or an aquifer. The aim of upscaling in
this case is to calculate a coarse-scale model which can reproduce the flow rates of the
different fluids. The coarse model should also provide a good approximation to the
saturation distribution in the reservoir with time.

The paths which the injected fluid takes through the reservoir depends on the forces
present:

©HERIOT-WATT UNIVERSITY December 2017 v2


31

 Viscous
 Capillary
 Gravity.

Therefore, the balance of forces should be taken into account during upscaling. Before
learning how to upscale two-phase flow, we show the effects which geological
heterogeneity may have on hydrocarbon recovery.

Consider the following simple model (Figure 26), with alternating horizontal layers of
100 mD and 10 mD (referred to as facies 1 and facies 2). We assume that the model is
filled with oil and connate water initially, and simulate a water flood, by injecting at
uniform rate at the left side, and producing from the right side (at constant bottom-
hole pressure). The density of the two fluids is the same for this example, so that there
are no gravity effects. In the first two examples, the relative permeabilities were the
same for the high and low permeability layers and the capillary pressure was zero.
Therefore the flood was viscous-dominated. In Case 1, the porosity was set constant
across the model (= 0.2). A diagram of the water saturation is shown in Figure 27a).
As expected the water moves quickest along the high permeability layers.

In Case 2, the porosity remained at 0.2 in the high permeability layers, but was changed
to 0.05 in the low permeability layers. This meant that the front advanced faster in the
low permeability layers, and so the flood front was more even.

Figure 26
Simple layered model for demonstrating viscous and capillary effects.

©HERIOT-WATT UNIVERSITY December 2017 v2


32

Figure 27
Water saturation distribution after in injection of 0.2 PV (pore volumes) of water: a)
Case 1 where the porosity was constant at 0.2, and b) Case 2 where the porosity was
0.2 in the high permeability layers and 0.05 in the low permeability layers.

A third case was run with using different relative permeabilities (with different connate
water saturations) and the capillary pressure was different in the different layers
(Figure 28). These curves are typical of a water-wet rock: the capillary pressure is much
higher in the low permeability facies, and the connate water saturation is higher. In
this case, water is imbibed along the low permeability layers, and also there is cross-
flow from the high permeability layers to the low permeability ones. Due to the effects
of capillary pressure, the front is nearly level in the two facies.

Figure 28
The relative permeabilities and capillary pressure used in Case 3.

©HERIOT-WATT UNIVERSITY December 2017 v2


33

Figure 29
Water saturation distribution, after the injection of 0.2 PV: Case 3, showing capillary
effects.

Figure 30 shows the recovery factor as a function of pore-volumes injected. All these
cases have the same permeabilities in the layers, and the same effective
permeability. However the total recovery depends on the capillary pressure and on
the porosity distribution.

Figure 30
Comparison of cumulative oil recovery in the 3 cases.

A summary of the models in Cases 1, 2 and 3 is given in Table 1.

Case Porosity Rel Perm/Pc Curve No Flow Regime


Facies 1 Facies 2 Facies 1 Facies 2
1 0.2 0.2 1 1 Viscous
2 0.2 0.05 1 1 Viscous
3 0.2 0.1 1 2 Viscous + capillary

Table 1
Properties for the first set of examples.

©HERIOT-WATT UNIVERSITY December 2017 v2


34

In the next example, graded models are used, as shown in Figure 31. There are two
versions: Case 4 with permeability increasing upwards (referred to as coarsening-up)
and Case 5 with permeability decreasing upwards (referred to as fining-up). The
permeabilities range from 200 mD to 1000 mD, the porosity was kept constant at 0.2,
and the same relative permeability curve was used. In this model, however, the
densities of the fluids were different: the density of water was set to 1000 kg/m3 and
that of oil was set to 200 kg/m3.

Figure 31
Coarsening-up and fining-up models for Cases 4 and 5.

Again a waterflood was performed, and the results are shown in Figure 32. Since water
is more dense than oil, water has a tendency to slump down. In Case 4 (coarsening-
up), this tendency is reduced by the fact that the viscous forces tend to move the fluid
faster in the upper layers. However, in Case 5 (fining-up), the slumping effect is
reinforced by the viscous force moving fluid faster along the lower layers. This means
that the breakthrough time is earlier in Case 5 than in Case 4, as shown in Figure 33.
The effective absolute permeability in these two models is the same, but the two-
phase flow effects are different, due to the effect of gravity. (If the density of water
equalled the density of oil, the recovery and watercut curves would be identical.)

Figure 32
Water saturation distributions for Cases 4 and 5.

©HERIOT-WATT UNIVERSITY December 2017 v2


35

Figure 33
Fractional recovery for the graded layer models.

In summary, in a viscous-dominated flood, permeability heterogeneity disperses the


flood front (Cases 1 and 2), so that breakthrough occurs earlier along high permeability
pathways, and the water cut curve rises less steeply. However, the effect of
heterogeneity also depends on the balance of fluid forces. In a water-wet system,
capillary pressure can help the front to advance more evenly along the layers (Case 3).
(More information on capillary effects is given in Section 6.1.) Gravity effects may
increase or reduce the viscous effects, depending on permeabilities in the model
(Cases 4 and 5).

4.2 Applying Single-Phase Upscaling to Two-Phase Problems


The previous section shows that the models with the same effective permeability can
produce different results for two-phase flow, and that these results also depend on the
balance of forces. Therefore, if we have two (or more!) phases in a reservoir, we should
take this into account when upscaling. Unfortunately, two-phase upscaling is difficult
– it is time consuming and often not robust. So many engineers only perform single-
phase upscaling.

©HERIOT-WATT UNIVERSITY December 2017 v2


36

Figure 34 (left side) shows a very heterogeneous 2D model, with correlated random
permeabilities. The permeability distribution was ln-normal (i.e. natural logs), with a
standard deviation of 2.0. The model is assumed to be in the horizontal plain. The
details of the model are given in Table 2. The model was upscaled using the pressure
solution method with no-flow boundaries. Three different scale-up factors were
and the coarse-scale models 1 and 3 are also shown in

Figure 34 (middle and right side).

©HERIOT-WATT UNIVERSITY December 2017 v2


37

Figure 34
Fine-and coarse-scale models used for demonstrating the effect of applying single-
phase upscaling to a two-phase problem.

Model Number of Cells Cell Dimension (m) Scale-up Factor


Fine 105 x 105 5 -
Coarse 1 21 x 21 25 5x5
Coarse 2 15 x 15 35 7x7
Coarse 3 7x7 75 15 x 15

Table 2
Details of models for check on single-phase upscaling for a two-phase system.

A waterflood with a quarter 5-spot well pattern was performed (i.e. 2 wells in
diagonally opposite corners). The same relative permeability curve was used for the
whole model (Figure 35), and the capillary pressure was set to zero. The flood was
therefore viscous-dominated. The viscosity of water was 0.3 and that of oil was 3.0.
The resulting recovery curves are shown in Figure 36. As one might expect, the error
increases with the scale-up factor.

The model was modified by reducing the standard deviation to 0.2 (lowering the
permeability contrast), and the simulations were repeated with the low heterogeneity
model. The recovery is shown in Figure 37 for the scale-up factor of 15 x 15. As
expected, the errors are smaller for the low heterogeneity model.

©HERIOT-WATT UNIVERSITY December 2017 v2


38

Figure 35
The relative permeability curve used for the random model. (Capillary pressure was
set to zero).

Figure 36
Comparison of recovery for different scale-up factors.

©HERIOT-WATT UNIVERSITY December 2017 v2


39

Figure 37
Comparison of recovery for models with different levels of heterogeneity (labelled
“hi” and “lo”).

To summarize, the use of single-phase upscaling in a two-phase system (with stochastic


permeabilities) may be accurate, if:
 the scale-up factor is small
 the permeability contrasts are small

From before, we know that the results will be affected by the balance of forces –
capillary pressure and gravity may adversely affect the results. In addition, upscaling
is more accurate when a flood is stable – i.e. when the injected fluid has a greater
viscosity than the produced fluid.

4.3 Improving Method of Single-Phase Upscaling


Commonly, engineers use non-uniform upscaling to improve the accuracy of single-
phase upscaling. For example, if we have a model with layers of different permeability,
it is best to upscale so that layers with similar permeability are grouped together. Then
the dispersion of the flood front will be similar in the fine and coarse models.

Consider a model with horizontal layers, as shown in Figure 38. There is a high
permeability streak running across the model. The model details are given in Table 3.
In both coarse-scale models there are 3 coarse cells in the vertical direction. In model
Coarse 1, the cells are each 5 m thick. However, in model Coarse 2, the thicknesses
are: 7 m, 1 m, 7 m, so that the high permeability streak is still resolved.

©HERIOT-WATT UNIVERSITY December 2017 v2


40

Figure 38
Model with high permeability streak.

Model No. of Cells Cell Size (m)


Fine 100 x 15 1x1
Coarse 1 20 x 3 5x5
Coarse 2 20 x 3 variable

Table 3
Details of models for example of non-uniform upscaling.

Figure 39 shows the recovery for these models. It can be seen that the model with
uniform coarse cells (Coarse 1) gives very inaccurate results, and the model which
maintains the high permeability streak (Coarse 2) is much more accurate.

Figure 39
Recovery for fine- and coarse-scale models.

In a practical upscaling application, much attention is paid to upgridding the model –


i.e. deciding how to amalgamate the layers, so that upscaling is as accurate as possible.
The coarse grid may also be non-uniform in the x- and y-directions.

©HERIOT-WATT UNIVERSITY December 2017 v2


41

There are some other approaches which makes single-phase upscaling more accurate
in a two-phase system. One is to improve the boundary conditions used when
performing numerical upscaling, and also to upscale the transmissibility between pairs
of coarse cells, rather than calculating the effective permeability (e.g. Zhang et al,
2005).

5. Two-Phase Upscaling Methods

5.1 Introduction to Two-Phase Upscaling


So far, when performing upscaling, we have assumed that there is only one phase
present, and that the flow is in a steady state. We only need to upscale the absolute
permeability. However, when there are two phases flowing, such as water displacing
oil, the system is not, in general, in a steady state. We need to simulate fine-scale
floods in order to upscale relative permeability and capillary pressure. This is referred
to as dynamic upscaling, and the upscaled relative permeabilities are known as pseudo
relative permeabilities, or pseudos. Pseudos can be calculated to take account of
physical dispersion, and also to compensate for numerical dispersion (Section 5.3.2).

When upscaling, we should use the phase permeabilities:

k f  kabs krf (30)

where “f” stands for fluid – oil, gas or water. Generally, we assume that both the
absolute and the relative permeabilities are homogeneous and isotropic at the
smallest scale ( krf , x  krf , y  krf , z ). As we upscale, the absolute and relative
permeabilities may become anisotropic ( krf , x  krf , y  krf , z ). To obtain effective (or
pseudo) relative permeabilities, the absolute permeability must be scaled-up
separately. Then the pseudo relative permeability is calculated as follows:

krf , x  k f , x kabs , x . (31)

Similar equations are used for flow in the y- and z-directions.

As mentioned above many engineers just use single-phase upscaling, even if the
reservoir they are trying to model is undergoing a water flood or gas flood. In many
cases, careful single-phase upscaling (with a non-uniform coarse grid and good choice
of boundary conditions) can be as accurate as two-phase upscaling. However, many
people have worked on methods for two-phase upscaling. In the next sections, we

©HERIOT-WATT UNIVERSITY December 2017 v2


42

give a brief description of a range of methods, and then focus on the best known
method (Kyte and Berry, 1975).

5.2 Steady-State Methods


The simplest method and quickest methods for upscaling are steady-state methods.
Although we stated above that two-phase floods are in general not in a steady state,
floods may sometimes come into a steady state, or quasi-steady state, where flows
and saturation are changing very slowly with time – i.e. Sw t  0 . Using this
assumptions, steady-state methods turn two-phase upscaling into a series of single-
phase upscaling calculations. So, once you can upscale for single-phase flow, you can
upscale two-phase flow using a steady-state method, with a little extra effort.

If we assume that we have an incompressible system and that Sw t  0 , then the
continuity equation (mass balance equation) for each fluid is given by:

q fx,in  q fz ,in  q fx,out  q fz ,out , (32)

where f = fluid. This equation is essentially the same as Equation (26), but we have two
– one for each fluid (oil and water). This leads to a set of pressure equations for each
fluid, similar to Equation (28), i.e.:
T x ,i 1/2, j  Tx ,i 1/2, j  Tx ,i , j 1/2  Tx ,i , j 1/2  Pi , j
Tx ,i 1/2, j Pi 1, j  Tx ,i 1/2, j Pi 1, j
(33)
Tx ,i , j 1/2 Pi , j 1  Tx ,i , j 1/2 Pi , j 1
0
where, in this case the transmissibilities include relative permeabilities (e.g.
Tx  kkrf yz x ).

There are several steady-state methods, depending on the balance of forces:


 capillary equilibrium
 vertical equilibrium (gravity-dominated flood)
 viscous-dominated steady-state.

We describe the capillary equilibrium method in a bit more detail below.

5.2.1 Capillary Equilibrium


Away from wells, the flow rate is very low – e.g. 1 m/day. If we assume that gravity
forces are negligible, oil and water may come into capillary equilibrium over small

©HERIOT-WATT UNIVERSITY December 2017 v2


43

distances (20 cm or less). This means that the saturation distribution is determined by
the capillary pressure curves.

The procedure for calculating the capillary-dominated relative permeability curves is


as follows:
1. Take a Pc level
2. Determine the water saturations for that Pc value, and then determine the
relative permeability values.
3. Calculate the average water saturation (pore-volume weighted, as shown in
Section 2.2.
4. Calculate the phase permeabilities: k f  kabs  krf
5. Calculate the effective phase permeability for each fluid
6. Calculate the absolute permeability
7. Calculate the relative permeabilities
8. Repeat the process with another value of Pc.

Figure 40 illustrates steps 1 and 2 of the process for a model which has two rock types
(any type of distribution). Steps 5 and 6 may be carried out analytically or numerically
depending on the complexity of the model.

Figure 40
Illustration of the Capillary Equilibrium method – determining the water saturation
and then the relative permeabilities for two different rock types.

5.3 Two-Phase Dynamic Upscaling


For dynamic (or non-steady-state methods), we need to perform a two-phase flow
simulation on a fine grid. There are a number of methods, and here we briefly describe
the Kyte and Berry (1975) method. An example of the steps in the calculation of Kyte
and Berry pseudos is given in the Appendix.

5.3.1 Outline of Kyte and Berry Method


This method is like an extension of the single-phase numerical upscaling, in that a fine-
scale simulation must be performed. In this case, a two-phase dynamic simulation is

©HERIOT-WATT UNIVERSITY December 2017 v2


44

required. The awkward thing, here, though, is deciding what boundary conditions to
use for upscaling in a reservoir model: the flow regime will depend on the location of
wells, and will be radial near a well. This is why two-phase dynamic upscaling is not
often performed. However, once a simulation has been performed, the basic idea is
to sum the flows at a coarse block boundary (on the down-stream side), calculate the
average pressure adjoining grid cells and use Darcy’s Law to determine an upscaled
phase permeability, and hence a pseudo relative permeability (Figure 41 and Equation
34).

Figure 41
General idea of the Kyte and Berry Method. The cell for which the pseudo is being
calculated is shaded in the darker colour.

Darcy’s Law between the two coarse blocks:

nz k f ,eff AP k f ,abs krf AP


Q f ,t   q f , k   (34)
j 1 X X

where j = 1, 2, ..nz is the cell counter in the z-direction, A is the cross-sectional area,
P is the difference in the pressure in the two coarse cells and X is the difference
between the cell centres. At the same time the average saturation needs to be
calculated for the cell. The calculations are repeated at intervals throughout the
simulation so that a table of relative permeability vs saturation can be constructed.

Equation (34) is a simplified form of the Kyte and Berry Method. A fuller derivation,
including gravity effects, is given in the Appendix.

5.3.2 Discussion on Numerical Dispersion


One advantage of pseudo-isation methods, such as that of Kyte and Berry is that they
can take account of numerical dispersion. (See Topic 3, Section 2.2.) When a
simulation is carried out using a larger grid, the front between the oil and water
becomes more spread out. However, the Kyte and Berry method counteracts this

©HERIOT-WATT UNIVERSITY December 2017 v2


45

effect by calculating the flows on the down-stream side of the coarse block, instead of
the middle. This is illustrated by a simple example. Figure 42 shows an example of
input relative permeability curves (“rock” curves).

Figure 42
Example of “rock” curves

If the water saturation is Sw = 0.5, the rock curves show that there is a small amount of
oil and water flowing. However, when the average saturation is 0.5 in the coarse block,
the distribution could be as shown in Figure 43.

Figure 43
Example of the water saturation in a coarse block

Since the water has reached only part-way across the coarse block, there should be no
water flowing out of the right side. The Kyte and Berry method calculates the pseudo
relative permeabilities using the flow on the downstream side of the coarse block, to
prevent water breaking through too soon. The pseudo water relative permeability
curve is moved to the right, relative to the rock curves, as shown in Figure 44.

©HERIOT-WATT UNIVERSITY December 2017 v2


46

Figure 44
Example of Kyte and Berry pseudo relative permeabilities.

5.3.3 Disadvantages of Dynamic Two-Phase Upscaling


There are several disadvantages of two-phase dynamic upscaling methods like Kyte
and Berry (1975). For example, the relative permeabilities may turn out to be negative
or infinite, so the pseudos need to be vetted before they can be used. Also, if you have,
say 10,000 grid cells in a coarse model, you could have 30,000 pseudos (one in each
direction). This is difficult to handle, so you need some way of grouping like pseudos
together.

Pseudos depend on a number of factors, so they are very case-specific:


 scale-up factor
 level of heterogeneity
 balance of forces (viscous, capillary gravity)
 proximity to wells.

Because of these problems full two-phase dynamic upscaling is rarely using for
upscaling.

Does the fact that reservoir engineers do not generally upscaling relative
permeabilities mean that reservoir simulation is being carried out carelessly? Not
necessarily – one has to consider all the other uncertainties in reservoir models, and
decide what is required. If we only need a coarse model to make a decision, that is
fine (Topic 1). The next section discusses the issue of what scale to start performing
simulations.

©HERIOT-WATT UNIVERSITY December 2017 v2


47

6. Scale Issues
By now, you know how to upscale models, either by averaging or using a numerical
method. However, one of the challenges of modelling and simulation is knowing how
to do this in practice. Firstly what scale should you start at? Should you start with fine-
scale models and upscale, or just make a coarse-scale model? The answer, as
mentioned in Topic 1, is that it depends on the decision being made. There is no need
to make a fine-scale model, if you can a make a decision based on a coarse-scale one.
Another issue is the amount of data available, because there is no point making a
detailed model if you have no data. Also, if there is a lot of uncertainty, you need test
many models to quantify the uncertainty and therefore do not have time to perform
flow simulations on models with millions of grid cells.

Some possible options for modelling are:


 Material balance model (one cell, “tank”) or a few cells, one for each compartment
of the reservoir. Negligible geological input – only some information about likely
volumes of compartments and transmissibilities between them. (See Topic 1,
Schiehallion Case Study.)
 Coarse-scale geological models – with some data from wells, upscaled to the block
size by averaging, without much thought about small-scale geological structures.
 Fine-scale geological models – as above, but requiring upscaling for flow
simulation.
 Detailed multi-scale models, where a lot of effort has gone in to making small-scale
models which account for representative sedimentary structures and upscaling
them.

Figure 3 is repeated here, as Figure 45 to remind you about scales. You may wonder
why anyone might bother to make models of sedimentary structures when these are
so small compared with the size of a reservoir. However, small-scale structures may
have a significant effect on two-phase flow, as shown below.

©HERIOT-WATT UNIVERSITY December 2017 v2


48

From Pickup and Hern (2002), adapted by Barkve (2004)


100

10 Seismic data Para-


sequences
Vertical thickness (m)

Log
1 Flow model

0.1 Core Geological model

Beds
0.01 Probe

Laminae
0.001
0.001 0.01 0.1 1.0 1 10 100 1000 10000
Horizontal length (m)

Figure 45
Length scales of sedimentary structures, data and models.

6.1 Effect of Small-Scale Structures


At small scales the flow is often capillary-dominated. Figure 29 (in Section 4.1) showed
a moderate capillary effect: the imbibition of water along the low permeability layer
made the flood front approximately level in the two layers (instead being ahead in the
high permeability layer, in the case of a viscous-dominated flood). In that model
(Figure 26), the layers were 1 m thick, and the grid cells were 10 cm square. If the size
of the model is reduced by a factor 100, so that the layers are 1 cm thick, and represent
sedimentary laminae, the effects of capillary pressure are much stronger, as shown in
Figure 46. In this case, strong capillary imbibition draws water into the low
permeability layers (black) so that the front advances faster in this layer

Figure 46
Example of capillary-dominated flood in a layered model.

©HERIOT-WATT UNIVERSITY December 2017 v2


49

Notice that, in Figure 46, there is little lateral variation in the shading, showing that the
water saturation is almost constant in each layer. This is because the front has been
spread out by the effects of capillary pressure, and the model is almost in capillary
equilibrium.

When the flow is across the layers, as in Figure 47, the effects of capillary pressure are
even more striking. This figure shows the same small-scale model of sedimentary
lamination, but rotated through 90o. When the injection rate is low (average frontal
advance rate of 0.3 m/day), the flood is capillary-dominated (Figure 47a), water (black)
has been imbibed into the low permeability layers leaving oil trapped in the high
permeability laminae (grey). As the injection rate is increased, the oil has more viscous
force and can overcome the capillary forces leading to less trapping of oil (Figure 47b).
In the case of Figure 47c, the flood is viscous dominated and all the movable oil has
been displaced by water.

Figure 47
Examples of across-layer flow. a) capillary dominated, b) intermediate, c) viscous-
dominated

The examples shown in Figure 46 and Figure 47 demonstrate the significance of


capillary effects at the small-scale. In some reservoirs small-scale structures can
therefore have a significant effect on oil recovery.

Note that we have assumed here that the rock is water-wet. The results would be
different if the rock was oil-wet, or intermediate-wet. (See Topic 7 for more
information on wettability.)

©HERIOT-WATT UNIVERSITY December 2017 v2


50

6.2 The Geopseudo Method


This is a method in which upscaling is carried out in stages using geological significant
length scales (Corbett et al, 1992; Ringrose et al, 1993). Usually models of typical
sedimentary structures (containing laminae) are created and upscaled to obtain the
effective properties at the bed scale. (If we assume that the beds are all about the
same size, this is upscaling to an REV.)

Models of typical sedimentary structures are created and permeability values are
assigned to the laminae (from probe permeameter measurements, or by analysing
core plug data). Relative permeabilities and capillary pressure curves are also assigned
to each lamina-type (e.g. by history matching SCAL experiments on core plugs, or from
pore-scale modelling (Topic 7)). Flow simulations are carried out to calculate the
effective single-phase permeability and the two-phase pseudo parameters. Additional
stages of modelling and upscaling may be required – e.g. upscaling from beds to bed-
sets.

Figure 48
Illustration of the Geopseudo Method.

Figure 49 shows an example of multi-level modelling and upscaling. This project


(described by Stephen et al, 2002) was carried out by a team of geologists and
engineers. The geologists split the rocks into 6 types – genetic units (GU). Three of the
genetic units are shown below – 1) clean sandstone, 2) argillaceous sandstone, and 3)
interbedded sandstone. Bed-scale models were generated for the argillaceous
sandstone and the interbedded sandstone, and these were upscaled by averaging
(geometric and arithmetic/harmonic methods, respectively). The relative
permeabilities were upscaled using the capillary equilibrium method (using the
appropriate averages). The next level of upscaling went from the bed scale (size 1m x
1m x 1.5m) to the scale of the blocks of the geological model (25m x 25m x 1.5m). The
Kyte and Berry (1975) dynamic upscaling method was used for this. One further level
of upscaling was performed to provide flow properties for the simulation model (cell

©HERIOT-WATT UNIVERSITY December 2017 v2


51

sizes 100m x 100m x 6m). This last stage of upscaling used flow-based single-phase
upscaling.

Figure 49
Example of multi-stage upscaling (SPE 78292, Stephen et al, 2002).

6.2.1 When to Use the Geopseudo Method


Geopseudo upscaling is time-consuming, and there is no point in upscaling from the
smallest scales, unless cores are available for the field. Cores must be studied to
identify the sedimentary structures present, and probe permeability measurements
should be taken to populate the small-scale models. Additionally, reliable SCAL data is
also required.

Ringrose et al. (1999) give a list of guidelines for when Geopseudo upscaling may be
necessary:

1) Are immiscible fluids flowing?


2) Are significant small-scale heterogeneities present? Specifically:
 Is the permeability contrast greater than 5:1?
 Is the layer thickness less than 20 cm?
 Is the mean permeability less than 500 mD?
3) What is the large-scale structure of the reservoir? In many cases, large-scale
connectivity may be the dominant issue, in which case, small-scale structure may
have to be ignored. The Weber and van Geuns (1990) classification may be used
to describe the large-scale structures:
 Layer cake reservoirs – small-scale structure will usually have primary
importance.
 Jigsaw puzzle reservoirs – small-scale structure may be important.

©HERIOT-WATT UNIVERSITY December 2017 v2


52

 Labyrinth reservoirs – small-scale structure will usually be of secondary


importance.

In the example described in Figure 49 and the accompanying text, the small-scale
heterogeneity did not actually have much of an effect in the base case model (after all
that effort!). In the base case where there was connected high permeability
sandstone, water from the aquifer did not need to flow through the lower permeability
heterogeneous structures. However, a sensitivity study was carried out with a layer of
interbedded sandstone across the model, so that water had to flow through this
structure. In this case, the geopseudo upscaling did make a difference.

7. Streamline Simulation
Finally, we describe a method which can be used to speed up fine-scale simulations in
order to avoid upscaling, namely streamline simulation. Here, we describe streamline
simulation qualitatively in a non-mathematical manner with reference to Figure 50
from the work of Gautier et al (1999). The basic procedure in streamline simulation
for a given permeability field (Figure 50(a)) is to calculate the pressure distribution by
solving a conventional pressure equation (see Topics 4 and 5). From this the iso-
potentials (pressure contours) can be calculated as shown in Figure 50(b); the gradient
of the pressures locally perpendicular to the iso-potentials are the streamlines as
shown in Figure 50(c). These streamlines are essentially the “paths” of the injected
fluid from the injectors (sources) to the producers (sinks). Since the velocity along
these paths is known (from Darcy’s Law using the calculated P ), we can work out
how far the saturation front moves along the streamline, l = v.t, where v is the (local)
velocity at that point on the streamline. Since v is known quite accurately, the advance
of the front along the streamline can be calculated accurately.

Note that the path along a streamline is one-dimensional, so the advance along a
streamline can be calculated using Buckley-Leverett theory (Topic 2), or can also be
calculated numerically. Also, streamline simulation can reduce the effect of numerical
dispersion and grid orientation (Topic 3). After we propagate the front along the
streamlines, the saturations will change over the reservoir model. These saturation
changes are then projected back onto the Cartesian grid as shown in Figure 50(d),
hence changing fluid mobilities. These updated mobilities can be used to recalculate
the pressures which, in turn, can be used to update the streamline pattern. This
process can be continued throughout the calculation. However, the calculation of the
pressure equation is what takes most computational time in most reservoir simulation,
so usually only a few pressure solves are performed.

©HERIOT-WATT UNIVERSITY December 2017 v2


53

Figure 50
The streamline method (Gautier et al, 1999).

The advantage of streamline simulation is that it is much quicker than conventional


simulation and, as mentioned above it can reduce some problems with conventional
simulation, such as numerical dispersion and the grid orientation effect. However, it
does this at the expense of losing some fluid physics. For example, using fewer
pressure solves means that the mobilities of the fluids are not updated so often. The
original streamline method did not take account of gravity (although this has been
rectified) and it ignores capillary pressure and compressibility.

MRST contains a module for calculating streamlines, and this forms an important
component of the MRST diagnostic tools. These tools are for rapidly assessing models
using single-phase incompressible flow. Once streamlines have been calculated, the
time of flight is computed. This is the time which “particles” of fluid take to travel from
an injector to a point in the model. Also, diagnostics can be used in complex models
to identify which regions are swept by a particular injector, or drained by a producer,
and which injectors are in communication with which producers. Figure 51 shows an
example of streamlines and time of flight for a model with two injectors and three

©HERIOT-WATT UNIVERSITY December 2017 v2


54

producers. This example is part of Lie (2016), Figure 13.1, and may be reproduced by
running the “showDiagnosticBasic” script in the /book/diagnostic folder.

Figure 51
Example of streamlines (white lines) and time of flight (shading) for a model with two
injectors and three producers (Lie, 2016). The colour bar on the right shows time of
flight in arbitrary units.

8. Concluding Remarks
This Topic has focussed on a subject which is essential for reservoir modelling and
simulation, namely how to calculate properties for a model. Firstly, a brief outline of
geological modelling was given. A full description of reservoir modelling is outside the
scope of this Course. However, the introduction to stochastic modelling and the
Tutorials demonstrate how to generate simple (Cartesian) models for flow simulation.

Most of the Topic concentrated on upscaling, which is a crucial part of the modelling
process (Ringrose and Bentley, 2015, P118). Data usually comes at a smaller scale than
is required for modelling, so some form of averaging is required. On the other hand,
some data may come from larger scales (e.g. well test data), and is already an effective
property. In which case, we need to understand how this data relates to any smaller-
scale measurements and to the scale of the model grid cell.

As stated above (and in Topic 1), there is no “correct” scale at which to make models:
it depends on the decision being made, and on the data available. At the appraisal
stage of reservoir development, many coarse-scale models may be tested, with no

©HERIOT-WATT UNIVERSITY December 2017 v2


55

explicit upscaling. On the other hand, for a mature field, a full Geopseudo approach
may be required.

9. References
Corbett, P.W.M., Ringrose, P.S., Jensen, J.L. and Sorbie, K.S., 1992. “Laminated Clastic
Reservoirs: The Interplay of capillary Pressure and Sedimentary Structure
Architecture”, SPE 24699, presented at the SPE Annual Technical Conference,
Washington, DC, 1992.

Gautier, Y., Blunt, M.J. and Christie, M.A., 1999. “Nested gridding and streamline-based
simulation for fast reservoir performance prediction”, Computational Geosciences, 3(3-
4), 295-320.

Kyte, J. R. and Berry, D. W., 1975. “New Pseudo Functions to Control Numerical
Dispersion”, SPEJ, August 1975, 269-276.

Lie, K-A, 2016. “An Introduction to Reservoir Simulation Using MATLAB: User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, www.sintef.no/mrst.

Nordahl, K. and Ringrose, P.S., 2008. “Identifying the Representative Elementary


Volume for Permeability in Heterolithic Deposits Using Numerical Rock Models”,
Mathematical Geoscience, 40, 753-771.

Ringrose, P. S., Sorbie, K. S., Corbett, P. W. M. and Jensen, J. L., 1993. “Immiscible Flow
Behaviour in Laminated and Cross-bedded Sandstones”, J. Petroleum Science and
Engineering, 9(2), 103-124.

Ringrose, P., Pickup, G.E., Jensen, J. and Forrester, M., 1999. “The Ardross Reservoir
Gridblock Analog: Sedimentology, Statistical Representivity and Flow Upscaling”, in R.
Schatzinger and J. Jordan, eds., Reservoir Characterization – Recent Advances, AAPG
Memoir 71, 265-276.

Ringrose, P. and Bentley, M., 2015. “Reservoir Model design: A Practitioner’s Guide”,
Springer, Dordrecht, Heidelberg, New York, London.

Stephen, K.D., Clard, J.D. and Pickup, G.E., 2002. “Modelling and Flow Simulations of a
North Sea Turbidite Reservoir: Sensitivities and Upscaling”, SPE 78292, presented at
the SPE 13th European Petroleum Conference, Aberdeen, Scotland, UK, 29-31 October,
2002.

©HERIOT-WATT UNIVERSITY December 2017 v2


56

Weber, K. J. and van Geuns, L. C., 1990. “Framework for Constructing Clastic Reservoir
Simulation models”. JPT, October 1990, p 1248 – 1297.

10. Appendix – The Kyte and Berry (1975) Method


A simple version of the Kyte and Berry (1975) method is presented here, using the grid
shown in Figure 52. The diagram shows two coarse grid blocks, each of which is made
up of 5 x 5 fine blocks. (We assume that the model is of constant thickness in the y-
direction.) The equations below show how to calculate the pseudo relative
permeabilities and capillary pressure for the left coarse block. It is assumed that a
water flood is being simulated and that the injector is somewhere to the left side of
the grid blocks shown, and the producer to the right.

Figure 52
Model used for describing the Kyte and Berry Method. The thickness of the model is
y.

The first step is to perform a fine-scale, two-phase simulation, saving the pressures and
inter-block flows at specified intervals of time. The method proceeds as follows:

1) Calculate the effective absolute permeability in the area shown in Figure 53, i.e.
half way between the two coarse blocks.

©HERIOT-WATT UNIVERSITY December 2017 v2


57

Figure 53
The area used for calculating the effective absolute permeability

Kyte and Berry approximate the effective permeability using the arithmetic average in
each column, and then taking the harmonic average of the columns. The area between
the two coarse blocks is used, for reasons explained below.

 z k
j 1
i ij

ki  (A1)
Z

where zj and Z are the thicknesses of the fine and coarse blocks, respectively. (In
this case, all the fine-scale blocks are of equal size.)

X
kI  7
(A2)
 x
i 3
i ki

where xi and X are the lengths of the fine and coarse blocks, and k I is the required
effective absolute permeability.

The pseudos are then calculated, at certain times during the simulation, as the water
saturations gradually increases.

2) Calculate the average water saturation. As you already know, this is a pore-volume
weighted average.

5 5

 S
j 1 i 1
 xi yz j
w,ij ij

Sw  5 5
(A3)
  x yz
j 1 i 1
ij i j

where ij is the porosity.

©HERIOT-WATT UNIVERSITY December 2017 v2


58

3) Calculate the total flow of oil and water out of the left coarse block (Figure 54).

5
q f   q f 5, j (A4)
j 1

where qf5,j is the flow of fluid “f” from fine block number (5,j).

Figure 54
Calculation of the total flow.

4) Calculate the average phase pressures in the central column of each coarse block.
In this example, we use the fine blocks in columns 3 and 8, the shaded areas in
Figure 55.

Figure 55
The cells used for averaging the phase pressures.

In the Kyte and Berry method, the pressures are weighted by the phase permeabilities
times the height of the cells (which in this case are all the same size). This is so that
more weight is given to regions where there is greater flow. However, there is no
scientific justification for using this weighting. In the first coarse block (numbered, I),
the average pressure is:

©HERIOT-WATT UNIVERSITY December 2017 v2


59

  
5

k
j 1
k
3 j rf 3 j z j Pf 3 j  g  j D3 j  D
P fI  5
(A5)
k
j 1
k
3 j rf 3 j z j

where D3j is the depth of cell (3,j) and D is the average depth of coarse cell I. The
 
term g  j D3 j  D is to normalise the pressure to the grid block centre. The average
pressure for coarse block II is calculated in the same manner, but using column 8
instead of column 3. The pressure difference is then calculated as:

Pf  P fI  P fII (A6)

5) The pseudo rel perms are then calculated using Darcy’s law. Firstly, calculate the
pseudo potential difference. (Potential is defined as   P   gz .)

 f  Pf  g D (A7)

where D is the depth difference between the two coarse grid centres. Then:

  q f X
k rf  (A8)
Z Y k I  f

6) Calculate the pseudo capillary pressure using:

Pc  PoI  P wI (A9)

Steps 2 to 6 are repeated, so that a table of krw  Sw  , kro  Sw  and Pc  Sw  is built up.

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 7 – Petrophysical and Fluid


Data for Reservoir Simulation
Contents
1. Introduction ............................................................................................................ 4
2. Petrophysical Data.................................................................................................. 4
2.1 Introduction to the Pore Scale ............................................................................. 4
2.2 Two-Phase Effects at the Pore Scale .................................................................... 7
2.2.1 Capillary Pressure .......................................................................................... 7
2.2.2 Drainage and Imbibition .............................................................................. 10
2.2.3 Drainage in a Capillary Bundle..................................................................... 11
2.2.4 Drainage in Connected Networks ................................................................ 14
2.3 Imbibition at the Pore Scale ............................................................................... 15
3. Laboratory Measurements ................................................................................... 19
3.1 Capillary Pressure Measurements ...................................................................... 19
3.2 Relative Permeability Measurements ................................................................ 19
4. Numerical Methods for Generating Curves ......................................................... 21
4.1 Methods for Capillary Pressure .......................................................................... 21
4.1.1 Rescaling of Mercury Injection Data ........................................................... 23
4.1.2 The Leverett J-Function ............................................................................... 24
4.1.3 Transition Zones .......................................................................................... 26
4.2 Methods for Relative Permeability..................................................................... 28
4.2.1 The Purcell Model........................................................................................ 28
4.2.2 The Burdine Model ...................................................................................... 30
4.2.3 The Brooks-Corey Model ............................................................................. 30
4.2.4 Network models .......................................................................................... 32
4.3 Hysteresis Phenomena ....................................................................................... 32
5. Wettability ............................................................................................................ 33
5.1 Introduction ........................................................................................................ 33

©HERIOT-WATT UNIVERSITY December 2017 v2


2

5.2 Pore-Scale Effects ............................................................................................... 35


5.3 Wettability Classification .................................................................................... 37
5.4 Impact of Wettability on Petrophysical Properties ............................................ 38
5.4.1 Effect of Wettability on Capillary Pressure.................................................. 38
5.4.2 Effective of Wettability on Relative Permeability ....................................... 39
6. Fluid Data.............................................................................................................. 40
6.1 Incompressible Fluids ......................................................................................... 41
6.2 Black Oil Simulators ............................................................................................ 42
6.3 Compositional Simulators................................................................................... 43
7. References ............................................................................................................ 44

©HERIOT-WATT UNIVERSITY December 2017 v2


3

Learning Objectives
After studying this topic, you should be able to:

 Describe how capillary pressure arises.


o State the Young-Laplace equation.

 Distinguish between drainage and imbibition.


o Outline the filling order for pores in a drainage process.
o Outline the filling order for pores in an imbibition process.

 Calculate the Pc for an oil/water system from the Pc for mercury/air.

 State the equation for the Leverett J-function and calculate Pc curves using the
Leverett J- function.

 Calculate relative permeabilities and capillary pressure using the Brooks-Corey


functions.

 Explain what is meant by wettability


o Describe how wettability affects Pc and rel perm curves

 Describe the fluid properties which are required for a reservoir simulator.
o Explain the difference between black oil simulation and compositional
simulation.

©HERIOT-WATT UNIVERSITY December 2017 v2


4

1. Introduction
In Topic 6 we discussed two subjects which are required for building reservoir models
for simulation, namely geological modelling and upscaling. We did not go into great
detail in modelling, since this is usually carried out by geoscientists. However,
engineers should always be involved in the construction of a geological model – e.g. in
deciding which structures will have important effects for flow, and which scale(s) to
use for simulation.

This Topic focuses other types of data which are required to set up a simulation,
namely relative permeability, capillary pressure and fluid PVT data. See the shaded
boxes in Figure 1 below.

Figure 1
Data required for reservoir simulation.

2. Petrophysical Data

2.1 Introduction to the Pore Scale


In reservoir simulation, as you know, we are dealing with grid cells, which have
homogeneous properties. But it should not be forgotten that, rocks are really made
up of grains with pores between them, and flow occurs through a network of
connected pores (Figure 2). In order to get a good understanding of properties such
as relative permeability and capillary pressure, we need to study what is happening at
the pore scale.

Figure 2 shows a CT scan of a rock sample of size 2mm x 2mm x 2mm. (There are 400
x 400 x 400 voxels in the model, and the voxels are approximately 5 m in length.) The

©HERIOT-WATT UNIVERSITY December 2017 v2


5

pore network is obviously very complicated, so it not surprising that reservoir


modelling usually starts at scales where the pore-scale effects have been averaged out.
Topic 6 introduced the concept of Representative Elementary Volume (REV). In fact,
the smallest scale for simulating flow in a porous medium is the scale of the first REV,
as shown in Figure 3. (This is the scale we would start at for the Geopseudo approach,
discussed in Topic 6.)

Figure 2
A close-up section of a rock (2 mm cube) and the extracted pore network. Left: CT
image, Middle: extraction of network; Right: 3D stochastic network model.

Figure 3
REV starting at the pore-scale. (From Lie, 2016, Figure 3.8, P55) Vv is the void volume
and Vr is the rock (grain) volume.

Although rocks are complex at the pore scale, many people have developed methods
for performing pore-scale modelling in order to link pore-scale structure with
macroscopic properties such as permeability and relative permeability, and to gain a
better understanding into how these properties arise. Early models considered the
pores as a bundle of parallel capillary tubes (i.e. narrow tubes). Gradually over the

©HERIOT-WATT UNIVERSITY December 2017 v2


6

years, models have become more sophisticated to represent a 3D network of pores.


One early example of a pore model is the Carman-Kozeny equation (Equation 1), which
relates the porosity and permeability. You do not have to memorise this equation, but
you may come across it in the MRST book (Lie, 2016). Carman and Kozeny assumed
that the travel times for a particle to flow through a tortuous pore network and an
equivalent homogenous porous medium are the same, and derived the following
equation:

3
k 2 (1)
2T 1    S S2
2

where T is tortuosity (i.e. the length of a tortuous flow path through the pore network
divided by the length of the system), and SS is the specific surface (i.e. the total surface
area of the grains divided by the volume of the sample). Unfortunately, it is difficult to
measure the parameters T and SS directly, although they can be calculated for simple
models (e.g. spherical grains).

Often empirical relationships are derived from core-plug measurements. Frequently,


it is assumed that there is a linear relationship between log(k) and . However, this is
only approximately true, and there is usually a large scatter in the data. The
relationship is different for different types of rock (Figure 4).

Figure 4
Example of the log(k)- relationship.

Over the past few decades, people have developed numerical pore network models at
the pore-scale models to predict properties. The input to these models is a network of

©HERIOT-WATT UNIVERSITY December 2017 v2


7

pores as shown in Figure 2 (right) and includes data on the distribution of pore radii
and their connectivity. The numerical models are similar to reservoir simulation
models in that they assume mass balance and then use the relationship between flow
rate and pressure in the pores, in order to construct a set of pressure equations.
However, in the case of fluid flowing through a pore, the Poiseuille Equation is used:

 r 4 P
Q (2)
8 L

where r is the radius of a pore. (You do not need to memorise this equation.) Pore
network models can be used for predicting permeability for single-phase flow, relative
permeabilities and capillary pressures for two-phase flow (and can also help to
understand the physics of three-phase flow).

2.2 Two-Phase Effects at the Pore Scale

2.2.1 Capillary Pressure


This section aims to give you a better understanding of capillary pressure which was
introduced in Course G19RA (Topic 7, Chapter 7). For many, the term “capillary
pressure” is a rather difficult concept to grasp. For example, Figure 5a) shows a narrow
tube (capillary tube) with oil and water. The stationary interface is not planar and Poil
> Pwater. If this is the case, why is there no flow of oil? Although capillary pressure may
seem problematic, we shall soon see that it is actually a very straightforward measure
to interpret.

Figure 5
a) Oil and Water in a capillary tube; b) the balloon analogy for capillary pressure.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

Let us first consider a rubber balloon (Figure 5b) that has been inflated to a certain
pressure (Pi) and then tied. (We will assume that this “experiment” is taking place in
pressure Po, usually atmospheric pressure). If a force balance is considered for one half
of the balloon (Figure 5c), then we can show that, at equilibrium, the elastic force
acting around the circular perimeter of the balloon must counterbalance the difference
in pressure projected onto the shaded cross section.

This leads to a relationship between the pressure difference across the balloon surface
and the radius of the balloon:
2
 Pi  Po   (3)
R

where  is the elastic tension characterising the balloon wall (dimensions of


Force/Length). Here is an example where a pressure difference exists between two
regions of fluid but no flow occurs — the elastic membrane of the balloon counteracts
this. Now, instead of thinking of a rubber balloon (where  is actually a function of R
itself), we can carry out a similar analysis for a gas bubble at pressure Pgas floating in
oil at pressure Poil. We can now write immediately:

2 go
P gas  Poil  
R
(4)

where go represents the interfacial tension between gas and oil. This relationship tells
us that the pressure difference across a small spherical bubble is larger than that across
a large spherical bubble.

Return now to the case where we have two fluids at equilibrium in a capillary tube
separated by a curved interface (Figure 6). This curved interface appears at the
microscopic scale when one of the fluids preferentially wets a solid surface. The
pressure difference across the fluid-fluid interface is known as the capillary pressure.
Note the difference between tube radius (rA) and the interface radius (RA) when the
contact angle is non-zero.

A little bit of trigonometry can be used to derive an equation for two fluids at
equilibrium in a circular cylinder, known as the Young-Laplace Equation. This relates
the pressure difference across a curved interface (i.e. capillary pressure Pc) in terms of
the associated contact angle, interfacial tension and pore (tube) radius.

rA
Note that RA  . (5)
cos 

©HERIOT-WATT UNIVERSITY December 2017 v2


9

Figure 6
Interface between two fluids in a capillary tube. The shaded fluid is the “wetting”
fluid.
(5)

Therefore:
 1 1  2 cos 
Pc   P gas  Poil    go    .
 RA RA  rA

And the Young-Laplace Equation is:

2 cos 
Pc  . (6)
rA

This is the Young-Laplace Equation. Although this example is for gas displacing oil, it
holds for any two fluids (e.g. oil and water).

The Young-Laplace Equation shows us that if a non-wetting phase is displacing a


wetting phase, more pressure is required for the non-wetting phase to invade smaller
tubes ( Pc 1 R ). Hence, it will invade larger tubes first.

In Figure 7a) there are 3 tubes of different radii, where R1 > R2 > R3. Let us assume
that the tubes are initially filled with water, which is the wetting phase, and oil is being
introduced to the system at gradually increasing pressures. In order to invade the
largest pore, the oil pressure must exceed the capillary entry pressure, i.e.

2 cos 
Po  Pc  .
R1

©HERIOT-WATT UNIVERSITY December 2017 v2


10

Figure 7
Invasion of 3 small tubes by non-wetting phase: a) filling of the tubes at different
stages; b) Pc vs wetting saturation.

Then, as the pressure is increased further, the oil will invade the medium-sized tube,
and then the smallest one. Figure 7b) shows capillary pressure vs Sw (wetting
saturation). To start with Sw = 1 – all tubes filled with water. Then it decreases and
the oil pressure increases, and the tubes are invaded successively.

Also, the Young-Laplace Equation (Equation 6) shows us that capillary pressure is more
important at small-scales, such as in a rock with microscopic pores.

2.2.2 Drainage and Imbibition


When considering two-phase flow, we use the term drainage to describe the invasion
of the non-wetting phase into the wetting phase, and the term imbibition to describe
the flow of the wetting phase into the non-wetting phase.

Therefore if the two phases are oil and water and the rock is “water-wet” (Section 5),
drainage means that oil is replacing water – the water saturation, Sw, decreases. An
example of drainage could be oil migration into a water-filled formation. If a water-
flood is being carried out in a water-wet reservoir, this is imbibition, and Sw will
increase. On the other hand, if the reservoir is “oil-wet”, the situation is reversed. For
example a water flood will be a drainage process. (Wettability will be discussed in
Section 5.)

It is known from lab experiments that relative permeability and capillary pressure
curves are different for the two processes. You need to study what is happening at the
pore scale to understand the processes.

©HERIOT-WATT UNIVERSITY December 2017 v2


11

2.2.3 Drainage in a Capillary Bundle


It is useful to first consider drainage in a capillary bundle (a set of parallel capillary
tubes). Assume a uniform distribution of pore radii, ranging from Rmin to Rmax. Also
assume that the system is water-wet and Sw = 100% initially, and that the contact
angle,  = 0 (cos() = 1). Oil cannot spontaneously invade the water-wet pores and
requires an increase in pressure for a displacement to occur.

Figure 8
Pore-size distribution for capillary bundle example.

This is a Primary Drainage process (i.e. there has been no oil in the system previously).
The steps in the drainage process and the corresponding drainage capillary pressures
are illustrated schematically in Figure 9. From the pore occupancies, we calculate the
water saturation Sw by summing the volume of the water-filled pores, divided by the
volume of all pores. Similarly, we may calculate the relative permeabilities (see below).

Step 1 – No oil can enter the system until the pressure is greater than the minimum
entry pressure, which is the capillary pressure corresponding to the tubes with the
largest radius. At this point:

2
Po  Pw  Pc1  .
Rmax

Step 2 – As the pressure is increased further, oil is able to enter smaller pores:
2
Po  Pw  Pc 2  ; Pc 2  Pc1 ; r2  Rmax .
r2

Figure 9b) shows that the oil saturation has increased and the water saturation
decreased.

Step 3 – The pressure is increased further so that all pores of radius r3 and greater have
been invaded.

©HERIOT-WATT UNIVERSITY December 2017 v2


12

2
Po  Pw  Pc 3  ; Pc3  Pc 2 ; r3  r2  Rmax .
r3

Figure 9
Steps in a primary drainage process, and the equivalent capillary pressure curve.

Step 4 (not shown) – Eventually, the pressure will be high enough to overcome the
entry pressure for the smallest pores. In this case:

©HERIOT-WATT UNIVERSITY December 2017 v2


13

2
Po  Pw  Pc max  .
Rmin

Also, in this example, the water saturation will go to zero. However, note that this is
only the case for our simple capillary bundle model.

As oil invades the system, the effective permeability to each phase will change – i.e.
the relative permeability will change, and shown in Figure 10.

Figure 10
Steps in a primary drainage process, and the equivalent relative permeability curves.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

Step 1 – the system is full of water, so krw = 1 and kro = 0.

Step 2 – The water saturation is Sw = Sw2 and krw falls quite rapidly since the water is
now flowing in smaller pores. Correspondingly, kro rises more rapidly since the oil is
flowing in the larger pores.

Step 3 – Again, the water saturation decreases, krw decreases and kro increases, as oil
invades more pores.

Step 4 (not shown) – finally, in this model the water saturation will reach zero: krw = 0
and kro = 1. However, as mentioned above, this is a simple system and this results
would not arise in a real system. (See next Section.)

2.2.4 Drainage in Connected Networks


There are some important differences between the capillary bundle model which was
used above, and a real inter-connected network of pores. Oil may be prevented from
entering a pore, even if it has enough pressure to overcome the capillary entry
pressure. One reason is accessibility – i.e. a large pore may be inaccessible because
there are smaller pores between the oil swept region and the pore in question (Figure
11 and Figure 12).

Figure 11 illustrates a uniform pore size distribution (p(r) = constant). (A1 + A2)
represents the pores which have a radius greater than r. Of these pores, A2 are
accessible to the non-wetting phase, and A1 are “shielded” by smaller pores, so are
inaccessible for the non-wetting phase. The accessibility factor may be written:

A2
Ar   .
 A1  A2 
(7)

Figure 11
Accessibility of large pores.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

Figure 12 shows a 2D pore network. This example shows a mercury/air system, but
the same process can apply to oil invading a water-filled, water-wet system. The
numbers beside the pores give their relative diameters.

a) The model is full of air.


b) As the pressure in the mercury is increased, the mercury will have enough pressure
to invade the widest pore.
c) The pressure has been increased further, so that the pore of diameter 10 can be
invaded. However, note that there are pores of diameter 11 and 12 in the model,
but they have not been invaded because access to them is blocked by narrower
pores.
d) The pressure is now high enough for mercury to enter pores with diameter 8 or
greater. At this stage there is a path of mercury from the inlet to the outlet. (This
is referred to as a percolating cluster of pores.)

Figure 12
Drainage in a 2D network.

2.3 Imbibition at the Pore Scale


We now consider imbibition, which is the opposite of drainage, i.e. in this case the
wetting phase displaces the non-wetting phase. Actually, imbibition is not just a

©HERIOT-WATT UNIVERSITY December 2017 v2


16

reversal of the drainage process, because of the pore-scale physics is different. In


drainage, oil displaced water by a piston-like displacement mechanism which was
governed by the Young-Laplace Equation. Imbibition at the pore scale can take place
by two distinct mechanisms – piston-like displacement and snap-off.

Piston-like displacement of oil by water is the reverse of drainage except that it occurs
when one of the phase pressures change such that Po - Pw < Pc = 2/r. The second
mechanism, snap-off, is associated with the flow of wetting phase (water) through
films. This film flow arises because, in a water-wet system, the water is in contact with
the grains, and may occupy irregular corners in pores. Figure 13 shows a schematic
diagram of a pore (with simple geometry for modelling purposes). The water can flow
in the corners, even though there is oil in the pore. Water films can swell around the
oil in a pore to form a “collar” which eventually - at an appropriate capillary pressure -
causes the oil to snap off thus occupying the space with water. This snap-off process is
shown schematically in Figure 14 for a single pore and in Figure 15 for a 2D
interconnected network. A given waterflood will generally consist of a mixture of the
two displacement mechanisms outlined above.

Figure 13
Schematic diagram of a pore with a triangular cross-section.

Figure 14
Illustration of snap-off. The wetting fluid is shaded and the non-wetting fluid is
white.

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Figure 15
The snap-off mechanism during imbibition in a 2D pore network.

The stages shown in Figure 15 are as follows:


a) Initially the network is full of oil (although it is water-wet).
b) The oil pressure has gradually decreased so that water can invade the pores. We
assume that film flow can occur along the sides of all the pores here. In that case,
snap-off will take place first in the narrowest pores, with diameters 1 and 2. At
this stage, there is still a percolating pathway for the oil through the wider pores.
c) As the oil pressure is reduced further, water is imbibed more, and snap-off occurs
in wider pores (with diameters of 6 or less). Note that some of the oil is trapped
in large pores on the left side, because it cannot escape through the narrower
pores which have become blocked with water.
d) Finally, there is a percolating pathway of water through the system. Oil is trapped
in 3 wide pores.

Figure 16 shows how the snap-off process relates to oil recovery. This figure shows a
“ganglion” of oil being snapped-off by water in a water-wet rock. The oil could only
escape through the connected cluster of oil filled pores (since there are no oil films in
a water-wet porous medium). We also note that the isolated blob of oil left behind in
this process in Figure 16 is “residual oil” since it is “trapped” and cannot now move
(unless viscous rather than capillary forces are invoked).

©HERIOT-WATT UNIVERSITY December 2017 v2


18

Figure 16
Schematic diagram of snap-off of a ganglion of oil.

It can be shown that the capillary pressure for snap-off is lower than that for piston-
like displacement. Indeed, in a strongly water-wet circular capillary (cos= 1), the
snap-off capillary pressure is approximately, Pc = /r i.e. half the value for piston like
displacement. Hence, if a capillary is occupied by oil and the oil pressure is lowered,
the capillary entry pressure for piston-like displacement will be reached first and – if
water is freely available, in adjacent pores for example – then piston-like displacement
will occur first. On the other hand, if the capillary entry pressure drops below the
piston-like entry pressure but no water front is available, then it will not fill with water.
However, if the oil phase pressure, Po, drops sufficiently (or Pw increases sufficiently)
that the snap-off capillary pressure is reached and there are water films to carry water
to that pore, then snap-off will occur. Hence, imbibition is a more complex process
than drainage and the balance between piston-like and snap-off events that occur
depends on a range of factors such as the range of pore sizes, pore-geometry (aspect

©HERIOT-WATT UNIVERSITY December 2017 v2


19

ratios), the connectivity of the network and the presence/absence of wetting films
(wettability).

Because the processes which take place in drainage and imbibition are different, this
means that relative permeabilities and capillary pressures are different for drainage
and imbibition. In fact, the properties depend on the saturation history. This is known
as hysteresis. We discuss this in more detail in Section 4.3

3. Laboratory Measurements
In this section, we will briefly mention a few different methods for measuring capillary
pressure and relative permeability. When setting up simulations it is important to
know how your data has been measured, so that you know if the data requires further
processing before you use it. You do not have to learn the details of this section. It is
included to make you aware of issues which you might have to address in the future
(e.g. in your project work for the BEng Program).

3.1 Capillary Pressure Measurements

Mercury Injection – although we are interested in oil-water capillary pressures, often


laboratory measurements are made by injecting mercury as the non-wetting phase.
Because of the difference in interfacial tension because air-mercury and oil-water,
these measurements have to be scaled before use (Section 4.1.2).

Porous Diaphragm – in this method oil invades a water-filled sample. This method
enables the residual water saturation to be measured. Occasionally water/air is used
to speed up the measurements, in which case the measurements have to be scaled
(Section 4.1.2).

Centrifuge Method – the sample is put into a centrifuge which spins at different speeds
providing different pressures to the oil phase.

Note that usually drainage curves are measured but, as mentioned above, imbibition
curves are different from drainage curves.

3.2 Relative Permeability Measurements

Steady-State Methods – in a steady-state method both oil and water are injected into
a sample simultaneously at a fixed ratio and at known flow rates. Steady-state
conditions are assumed to be reached once the inlet and outlet fluxes of each phase
have equilibrated and/or a constant pressure drop is seen across the sample. (This
may take a long time.) Once equilibration has been reached, Darcy’s Law can be used
for each phase in turn, resulting in a pair of relative permeability values valid for that
particular saturation. The fluid flux ratio is then changed, whilst keeping to total flow

©HERIOT-WATT UNIVERSITY December 2017 v2


20

rate constant, yielding a second set of data once a new steady-state has been achieved.
This procedure can be repeated for a number of flux ratios, thereby building up a set
of relative permeability curves which span the entire saturation range. Although the
time involved in extracting such data is clearly an important concern, steady-state
measurements performed at low rates should be considered most indicative of
reservoir behaviour. There are a number of different steady-state methods, which
we will not discuss. An example of some steady-state relative permeability
measurements is shown in Figure 17.

Figure 17
Oil-Water relative permeabilities for a Berea Sandstone sample, measured by
different steady-state techniques.

Unsteady-State Methods – In this case one phase is injected into a core and displaces
another. This method is faster than the steady-state method. However, the results
need more work to interpret. The injection rate is made high enough so that capillary
effects are negligible, and fractional flow theory can be used to calculate the relative
permeabilities. (See Topic 2 on Buckley-Leverett Theory.) The actual calculations are
a bit more complex and use a method developed by Johnson Bossler and Naumann
(1959), commonly referred to as the JBN method.

The disadvantage of this method is that the relative permeabilities are only obtained
after breakthrough of the injected phase. See Figure 18a). Sometimes the remainder
of the curves are extrapolated, so they look like complete relative permeability curves,
but they are not (Figure 18b).

©HERIOT-WATT UNIVERSITY December 2017 v2


21

Figure 18
Example of the results of an unsteady-state relative permeability measurement: a)
results from measurements; b) with the rest of the curves extrapolated.

Obviously it is useful to know whether a steady-state or an unsteady-state method has


been used to measure relative permeabilities. This will be stated on the laboratory
report. Steady-state measurements are more desirable than unsteady-state ones,
particularly since flow rates in a reservoir are slow and the fluids are probably in a
steady-state over short distances (10 – 20 cm).

4. Numerical Methods for Generating Curves


We often have to undertake a simulation study without possessing all of the data we
would ideally require — perhaps we have only one experimental capillary pressure
curve at our disposal and perhaps it comes from a part of the reservoir not directly
related to our current study. We need some way of inferring reasonable data from
those available to us at the time and, although we may have to make some gross
assumptions, we should be able to invoke our knowledge of flow at the pore-scale to
help us.

4.1 Methods for Capillary Pressure


Let us begin by reminding ourselves of the drainage case (drainage curves are
frequently used in simulators — often erroneously). Remember that, in drainage, we
are increasing the pressure difference between the non-wetting and wetting phases.
As Pc increases, the radius of interface curvature decreases and, using a capillary tube
analogue, the Young-Laplace equation tells us that the non-wetting phase begins to

©HERIOT-WATT UNIVERSITY December 2017 v2


22

invade the porous medium when Pc  2 cos   Rmax — i.e. at the so-called
displacement pressure of the sample. As Pc continues to increase, the non-wetting
phase invades progressively smaller pores. Hence, the non-wetting phase saturation
gradually increases with Pc. The resulting plot of Pc vs Sw is the drainage capillary
pressure curve (Figure 19).

Figure 19
Example of a Pc drainage curve

We can now use this discussion to infer the type of capillary pressure curve that may
result from different samples. For example, how would the pore size distribution affect
the curve? We already know that the displacement pressure is affected by Rmax — the
largest pore in the sample — but it should also be clear that the range of pore sizes in
a sample can greatly affect the shape of the corresponding Pc curve (Figure 20). Flat
plateaux indicate samples that are fairly homogeneous at the pore scale, whilst steeper
curves indicate a large variance in the distribution. Moreover, fine textured rocks with
small cemented grains can be expected to exhibit higher capillary pressures at a given
saturation than coarse-textured media, and also higher displacement pressures. So,
from our basic understanding of capillary pressure at the microscopic scale, we have
been able to infer a great deal about how we would expect capillary pressure curves
to vary at the continuum (macroscopic) scale.

In Figure 20, the curves all have the same displacement pressure, so Rmax is the same
in each case. The curves labelled 1, 2 and 3 have progressively less well sorted grains,
so the pore sizes are more varied, and the size of the smallest pore decreases.

©HERIOT-WATT UNIVERSITY December 2017 v2


23

Figure 20
Pc curves for rocks with different amounts of fine-scale grains.

4.1.1 Rescaling of Mercury Injection Data


One of the most common ways to derive oil-water or gas-oil capillary pressure curves
is to rescale mercury injection data (which is routinely measured and relatively cheap
to obtain). Consider mercury injection into a porous medium containing a large
exterior pore of radius Rmax. The capillary pressure required for this pore to be invaded
by mercury is given by the Young-Laplace equation:

2  cos  mercury / air


 Pce mercury / air  (8)
Rmax

where Pce is the capillary entry pressure.

If we now consider the same medium, filled with water, undergoing oil injection, then
the requisite oil-water capillary pressure is now:

2  cos  oil / water


 Pce oil / water  (9)
Rmax

©HERIOT-WATT UNIVERSITY December 2017 v2


24

Now, eliminating Rmax from Equations (8) and (9) we arrive at the scaling relationship:

 cos  oil / water


 Pce oil / water   Pce mercury / air . (10)
 cos  mercury / air

This clearly holds for any Pc-value, and so a complete oil-water capillary pressure curve
can be obtained from the mercury injection data by applying Equation (10) at each
saturation.

Experience shows that the ratio of interfacial tensions and cosines on the right hand
side of (10) (often difficult to measure accurately) should have a value of approximately
1/6. However, different ratios have sometimes been needed to reconcile experimental
mercury-air and oil-water data from different rock-types (Figure 21 — the ratio is
referred to as—“Factor” in these plots). Such differences may be due to interactions
between contact angle and pore geometry.

Figure 21
Mercury air Pc curves (axis on right side of graphs) scaled to water/gas curves (axis on
left side of graphs).

4.1.2 The Leverett J-Function


Pc curves are obviously affected by the pore-size distribution, the interfacial tension
the contact angle and the pore structure. Leverett (1941) developed a dimensionless
group – called a J-function – to account for these effects. The J-function has become

©HERIOT-WATT UNIVERSITY December 2017 v2


25

an important tool for capillary pressure interpolation and extrapolation and increases
the utility of a single capillary pressure curve. It allows us to adapt a single data set to
other areas of a reservoir where data may be unavailable.

It can be shown using capillary bundle models that a “representative pore radius” can
be estimated as follows:
k
R . (11)

Hence a dimensionless group can be formed by defining:

1/2
Pc k
J   (12)
 cos   

This can be applied at a number of different saturation values — at each saturation,


Pc(Sw) is determined and Equation (12) used to find J(Sw). Often, the contact angle is
difficult to ascertain with any accuracy, and the cos term is often set to unity.

Having found J(Sw), capillary pressure curves corresponding to a range of different


permeability and porosity values (and different fluid combinations via the cos term)
can be determined by simply inverting the J-function, as follows:

J  S w   cos 
Pc  S w   (13)
k  
1/2

Example J-functions are shown in Figure 22 together with a derived set of Pc-curves.

©HERIOT-WATT UNIVERSITY December 2017 v2


26

Figure 22
Left: One J-Function derived from Pc measurements of different cores. Right: Pc curves
for rocks with different permeabilities calculate from one J-function.

4.1.3 Transition Zones


We conclude this section on capillary pressure curves with a brief discussion of
transition zones. In reservoirs, there is not an abrupt change at the interface between
gas, oil and water. Instead there is a transition zone, the size of which depends on the
capillary pressure.

In Figure 23 (right), we see that the hydrostatic oil and water gradients meet at the
OWC (Po=Pw). Above this contact, Po>Pw and capillary pressure increases as we go
further up the reservoir.

Note that, actually Pc = 0 (i.e. Po = Pw) at the free water level (FWL). The oil water
contact (OWC) is slightly above this, and is at the depth where the oil saturation starts
to increase. However, sometimes this difference is overlooked.

©HERIOT-WATT UNIVERSITY December 2017 v2


27

Figure 23
Capillary Pressure and transition zones.

Capillary pressure data can also be used to infer fluid saturations as a function of depth
within a transition zone. Assuming that oil and water pressures remain continuous
throughout the entire height of the zone, we can write equations for each hydraulic
gradient (z measures height above OWC (FWL) and increases vertically upwards):

dPw
  g w (14)
dz

dPo
  g o (15)
dz

Subtracting (14) from (15) leads to:

dPo dPw dPc


   g   w  o  (16)
dz dz dz

which can be integrated to give:

Pc  g  w  o  z . (17)

Hence, any capillary pressure curve can be re-plotted to give saturation with depth
information. An example is shown in Figure 24.

©HERIOT-WATT UNIVERSITY December 2017 v2


28

logs

Approximate oil-
water contact

Figure 24
Example of the variation of water saturation with depth.

4.2 Methods for Relative Permeability


Multiphase flow experiments are costly, difficult and time-consuming to carry out and
it is often infeasible to perform a large number of laboratory sensitivity studies. It
would therefore be highly desirable if a cheap, predictive model for relative
permeability could be developed that only required cheap, easily-accessible data as
inputs. In fact, a number of models are already currently available and we will discuss
three such approaches below.

4.2.1 The Purcell Model


The Purcell model seeks to utilise cheap capillary pressure data in order to predict
more expensive relative permeabilities. The method was derived using the capillary

©HERIOT-WATT UNIVERSITY December 2017 v2


29

bundle model. Here we simply present an outline of how to use the method. You do
not have to learn this method, or the Burdine version, described below. In the steps
below the subscript “w” is for the wetting phase and “nw” for the non-wetting phase.

(i) Obtain capillary pressure data Pc(Sw)

(ii) Plot the curve 1 Pc2 vs Sw (See Figure 24)

(iii) Find the area under the entire curve (= kabs)

(iv) For any chosen value of water saturation (Swt, say), calculate the area under the
1 Pc2 curve from 0 to Swt. This gives kwt

(v) For the same value of water saturation, calculate the area under the 1 Pc2 curve
from Swt to 1. This gives knwt

(vi) Use the values obtained from (iii) – (v) to determine relative permeabilities

Figure 46
Curves used for Purcell method of determining relative permeability.

The relative permeabilities are then calculated using the formulae:

©HERIOT-WATT UNIVERSITY December 2017 v2


30

S  Swt
k S   dS  Pc 
2

krw  S wt   w wt  S 0
S 1
, (18a)
 dS  Pc 
2
kabs
S 0

S 1
 dS  Pc 
2
k S 
krnw  S wt   nw wt 
S  S wt
S 1
. (18b)
 dS  Pc 
2
kabs
S 0

There is one clear drawback using this approach: the relative permeabilities add up to
unity over the full saturation range and are therefore completely symmetrical. This
limitation led Burdine to add the following improvement.

4.2.2 The Burdine Model


The Burdine model simply takes the Purcell model and re-scales the endpoints through
the factors:

Sw  Swi 1  Sw  Sor
w  Sw   nw  Sw   (18)
1  Swi  Sor 1  Swi  Sor

(This is assuming that the wetting phase is water and the non-wetting phase is oil. Swi
is initial water saturation, or connate water saturation, Swc.)

The curves are therefore defined over the wetting-phase saturation range Swi to (1 -
Sor). Although a number of gross assumptions have been made along the way, the
Burdine method offers a cheap alternative to laboratory measurement of relative
permeabilities. It is easily coded into a spreadsheet and is often carried out by
practicing reservoir engineers — experience appears to show that the wetting phase
prediction is often fairly good, whilst the non-wetting curve is less well reproduced.

4.2.3 The Brooks-Corey Model


Although the models described above yield relative permeability curves that are
qualitatively reasonable, reproduction of experimental data is only achieved via the
introduction of some form of empiricism. Many studies have subsequently relied
entirely upon empirical curve fitting techniques. One of the most popular empirical
correlations is that due to Corey (1954), who proposed the following:

krw  Seff4 (19)

©HERIOT-WATT UNIVERSITY December 2017 v2


31

krnw  1  Seff  1  S 
2 2
eff (20)

The effective, or normalised saturation goes from 0 to 1, and is defined as:

S w  S wc
Seff  . (21)
1  S wc  Sor

However, it was soon discovered (not surprisingly) that the exponents in Corey’s
original equations would have to be varied in order to fit different materials. They
were consequently generalised by Brooks and Corey (1964) to:

 2  3  
krw  Seff (22)


krnw  1  Seff  1  Seff 2   
2
(23)

Seff   Pcb Pc   Pc  Pcb 



(24)

with Pcb representing the breakthrough capillary pressure and  the “pore size
distribution index”. Both of these parameters have to be determined experimentally:
Seff vs Pc is plotted on a log-log scale and a straight line fitted, the slope gives  and the
intercept with Seff =1 is assumed to give Pcb.

Simpler Corey Model


Many engineers use simpler formulae for relative permeability – and refer to the
relative permeabilities as a Corey model.

w
krw  krwm Seff (25)

 
 nw
krnw  krnwm 1  Seff (26)

Where w and nw are exponents for the wetting and non-wetting fluids, and krwm and
krnwm are the maximum values of the relative permeabilities.

©HERIOT-WATT UNIVERSITY December 2017 v2


32

4.2.4 Network models


Numerical pore network models were mentioned briefly in Section 2.1. These can
simulate a range of two-phase processes occurring at the pore-scale. A comparison
between experiment and network model results is shown in Figure 25. There is good
agreement between simulation and experiment.

Figure 25
Left: Comparison of network model relative permeabilities with experiment. Right:
Inverse capillary pressure data.

4.3 Hysteresis Phenomena


We have already mentioned that both capillary pressure and relative permeability
curves exhibit hysteresis — that is, they depend upon saturation history (Section 2.3).
There are a number of possible causes for hysteresis but the three main effects are the
following:

(i) Contact angle hysteresis — advancing and receding contact angles differ
(reflected in Pc through the Young-Laplace equation);

(ii) Pore structure hysteresis — sloping pore walls mean that pores fill and empty at
different capillary pressures;

(iii) Topological hysteresis — imbibition and drainage processes are different


topologically (film-flow versus fingered invasion).

There is often such a difference between imbibition and drainage curves (both capillary
pressure and relative permeability) that we must make sure that we are using the
correct set of curves in our reservoir simulation studies. (For instance, we should not
be using drainage curves as simulation input when modelling a waterflood in a water-

©HERIOT-WATT UNIVERSITY December 2017 v2


33

wet reservoir). Built-in numerical models are available in Eclipse and other commercial
reservoir simulators to account for flow reversals at intermediate saturations.
Hysteresis is not, however, included in the codes provided in MRST.

Figure 26
Example of hysteresis in capillary pressure and relative permeability.

5. Wettability

5.1 Introduction
The term wettability refers to the wetting preference of a solid substrate in the
presence of different fluid combinations (liquids and/or gases). Figure 27 shows how
the contact angle changes depending on the fluids, and the solid surface. The wetting
phase is defined as the fluid that contacts the solid surface at an angle less than 90o.

©HERIOT-WATT UNIVERSITY December 2017 v2


34

Figure 27
Examples of contact angles for different fluids and different surfaces.

This figure clearly demonstrates an important issue, namely, that the wetting
preference of a rock depends not only upon the fluids involved but also upon the
mineralogy of the rock surface. For instance, we see that, in the presence of
isoquinoline, water is non-wetting on a silica substrate but wetting on a calcite
substrate. In cases when a solid has no wetting preference, (i.e. when the angle
separating the two fluid interfaces is close to 90o) the system is said to be of “neutral”
wettability.

The importance of rock wettability cannot be over-emphasised, as it affects almost all


types of core analysis: capillary pressure, relative permeability, waterflood behaviour,
and electrical properties (see Anderson, 1987, for an excellent series of review
articles). The simple reason for this is that wettability affects the location, flow, and
distribution of fluids in a porous medium — hence, most measured petrophysical
properties must be affected.

In most of what has been described in this chapter, the assumption has been made
that the system under consideration was water-wet. Indeed, historically, all reservoirs
were believed to be strongly water-wet and almost all clean sedimentary rocks are in
a water-wet condition. An additional argument for the validity of the water-wet
assumption was the following: the majority of reservoirs were deposited in an aqueous
environment, with oil only migrating at a later time. The rock surfaces were
consequently in constant contact with water and no wettability alterations were
possible as connate water would prevent oil contacting the rock surfaces. However,
Nutting (1934) realised that some producing reservoirs were, in fact, oil-wet (the rock
surface was preferentially wetted by oil in the presence of water) and it is now

©HERIOT-WATT UNIVERSITY December 2017 v2


35

generally accepted that water-wet reservoirs are the exception rather than the rule.
Table 1 shows wettabilities for a sample of 55 reservoirs. In this sample half of the
silicate reservoirs were oil-wet and nearly all of the carbonate reservoirs.

Contact
Silicate Carbonate Total
Angle
Reservoirs Reservoirs Reservoirs
(degrees)
Water-wet 0 to 75 13 2 15
Intermediate-wet 75 to 105 2 1 3
Oil-wet 105 to 180 15 22 37
Total 30 25 55

Table 1
Wettability of a selection of reservoirs.

We now know that wettability is a rather complicated issue and that there are a
number of different factors affecting reservoir wettability, including: surface-active
compounds in the crude oil, brine composition, mineralogy and pressure and
temperature. What’s more the wettability may vary throughout a reservoir.

Having discovered that most reservoirs are not water-wet, we now have to modify
everything that we have learned so far — pore-scale physics, capillary pressure models,
relative permeability models, and network models. However, this is not as difficult as
it may first appear; we already understand drainage and imbibition processes at the
pore scale, we know how to define capillary pressure, and we appreciate the dynamics
underlying relative permeability measurements. The following discussion should help
you apply your previous water-wet knowledge to systems that are not water-wet.

5.2 Pore-Scale Effects


The effect of wettability at the pore-scale is shown in Figure 28. If a rock is water-wet,
we have already seen that there is a tendency for water to reside in the tighter pores
and to form a film over the grain surfaces. Oil (the non-wetting phase) resides in the
larger pores. In this case, the term “imbibition” – a process whereby a wetting phase
displaces a non-wetting phase – would refer to the displacement of oil by water. The
term “drainage” would apply to oil displacing water. In an oil-wet system, however,
the situation is reversed – oil now forms a thin film over the grain surfaces and water
fills the larger pores. Consequently, in an oil-wet medium, “imbibition” refers to the
displacement of water by oil, whilst “drainage” refers to the displacement of oil by
water. These differences clearly have major implications for waterflooding reservoirs.
If a reservoir is oil-wet, a waterflood is a drainage displacement, oil is flushed from
small pores, and oil production via film-flow also becomes important.

©HERIOT-WATT UNIVERSITY December 2017 v2


36

Figure 28
Comparison of a water flood in a) water-wet and b) oil-wet rocks.

The foregoing discussion has clarified some of the pore-scale physics affecting water-
wet and oil-wet media. However, given the complex nature of wettability alterations
in reality, a more realistic representation of wettability at the pore scale may be that
shown in Figure 29. Here, a combination of water-wet and oil-wet pathways exists that
allows film-flow access to the displacing phase and film-flow escape for the displaced
phase. Hence, a waterflood would now consist of a combination of imbibition and
drainage events: water initially imbibing along water-wet pathways (displacing oil from
small pores) and then requiring an overpressure to drain oil from the larger oil-wet
pores. It is clear that the underlying mineralogy and surface chemistry of the system
would determine the connectivity and topology of the “wettability pathways” but how
can we determine these pathways – in short, how can we quantify the wettability of a
porous medium?

©HERIOT-WATT UNIVERSITY December 2017 v2


37

Figure 29
Network of water-wet and oil-wet pores.

5.3 Wettability Classification


Wettability can be classed as being uniform or non-uniform as follows:

Uniform – the wettability of the entire pores pace is the same (100% water-wet, 100%
oil-wet, or 100% “intermediate-wet”) and the contact angle is essentially the same in
every pore;

Non-Uniform – this is more characteristic of hydrocarbon reservoirs. The pore space


exhibits “heterogeneous” wettability, with variations in wetting from pore to pore (and
possibly within a pore) – say 70% water-wet pores and 30% oil-wet pores. We can
introduce 2 subdivisions (
Figure 30):

Mixed-wet – a certain fraction of the largest pores are oil-wet (there are valid
depositional arguments for how this may come about).

Fractionally-wet – no size preference for oil-wetness (there are valid mineralogical


arguments for this).

©HERIOT-WATT UNIVERSITY December 2017 v2


38

Figure 30
Pore size distributions, indicating the difference between mixed wet and fractionally
wet.

Given our knowledge of pore-scale displacements, it is clear that each type of


wettability distribution will yield different capillary pressures and relative
permeabilities (and recoveries), and we should be aware of this when we come to
interpret SCAL and wettability test data.

5.4 Impact of Wettability on Petrophysical Properties

5.4.1 Effect of Wettability on Capillary Pressure


In oil-water systems, the capillary pressure, by convention, is taken as the oil pressure
minus the water pressure. Therefore, in a water-wet system, Pc is positive. However,
in an oil-wet system, Pc will be negative – i.e. in a drainage process, the water pressure
must be greater than the oil pressure for water to enter the system.

If the wettability is heterogeneous, then the Pc curve will have both positive and
negative parts. When the pore network contains a mixture of water-wet and oil-wet
pores, a waterflood initially proceeds as an imbibition, with water spontaneously
imbibing along water-wet pathways displacing oil from the smallest pores. Eventually,
however, the water has to be forced into oil-wet pores and the displacement becomes
one of drainage.

©HERIOT-WATT UNIVERSITY December 2017 v2


39

Figure 31
Example of a capillary pressure curve for an intermediate-wet rock sample. 1)
drainage, 2) spontaneous imbibition, 3) forced imbibition (or drainage of water).

5.4.2 Effective of Wettability on Relative Permeability


As wettability controls the distribution of phases within the pore space, it is hardly
surprising that wettability has a major impact upon relative permeability curves and
subsequent reservoir performance. We can use our pore-scale modelling knowledge
to help explain the differences seen between the slopes of water-wet and oil-wet
curves (
Figure 32). In the water-wet case, water imbibes via the smallest pores in the system,
leaving oil resident in large, fast-flowing pores. Consequently, we would expect the oil
relative permeability curve to decrease slowly with increasing water saturation and the
water relative permeability endpoint to remain low – this is exactly what is observed.
Conversely, waterflooding in an oil-wet medium should lead to a rapid decrease in oil
relative permeability, together with a high water endpoint – once again, this is what is
observed.

©HERIOT-WATT UNIVERSITY December 2017 v2


40

Figure 32
An examples relative permeabilities for a water-wet and an oil-wet system.

In fact, the key features of such curves were presented by Craig (1971) who indicated
the differences between the two in the form of several rules of thumb (see
Table 2). Many subsequent experimental studies have agreed with these ideas.

Water-Wet Oil-Wet
Connate water saturation Usually greater than 20 Generally less than 15%
to 25%
Saturation at which oil Greater than Sw = 50% Less than Sw = 50%
and water relative
permeabilities are equal
Relative permeability to Generally less than 30% Greater than 50%, and
water at the maximum approaching 100%
water saturation

Table 2
Craig’s Rules of Thumb

6. Fluid Data
In this Topic, we have considered petrophysical properties in great detail, because they
determine the distribution of fluids in the reservoir (at the pore-scale), and how they
flow. However, the distribution of fluid in a reservoir also depends on density, and
fluid flow depends on both density and viscosity. So, before leaving this topic, we
briefly review fluid properties.

©HERIOT-WATT UNIVERSITY December 2017 v2


41

In Topics 4 and 5 we derived the flow equations and showed how to discretise them
for numerical flow simulation. The coefficients in the pressure equations depend on
the density and viscosity, which in turn depend on pressure (so they are non-linear
equations). We therefore need to input tables of density and viscosity vs pressure, or
equations to enable these properties to be calculated.

6.1 Incompressible Fluids


To simplify some simulations, MRST has a solver which assumes that fluids are
incompressible (Lie, 2016). In this case a minimum amount of data is required for each
fluid. Firstly we need to know the viscosity of the fluid, because this determines the
flow velocity, given the pressure gradient and permeability, i.e. viscosity is required in
Darcy’s Law:

k P
u (27)
 x

where u is the Darcy velocity.

If there is only one incompressible fluid flowing in the model, this is all the fluid data
which is required, because in an incompressible system, the density is constant and
therefore cancels in the mass balance equation – e.g. Equation (10) in Topic 4, which
is reproduced here as Equation (28):

     u 
 (28)
t x

However, if there are two (or more) fluids flowing, the density is also required.
Equations (29) and (30) are the same as Equations (55) and (56) in Topic 4, and give the
general form of the mass balance equation for two-phase flow.

   So    kkro  Po  z  
       o g   . (29)
t  Bo  x  o Bo  x  x  

   Sw    kkrw  Pw  z  
  
t  Bw  x  w Bw  x    w g x   . (30)
  

Whilst deriving these equations in Topic 4, we used    sc B , where  is the density


in the reservoir, sc is the density at the surface (standard, or stock tank conditions),

©HERIOT-WATT UNIVERSITY December 2017 v2


42

and B is the formation volume factor (units of rb/stb, or rm3/sm3). Then sc cancelled
in the mass balance equation. When using the incompressible routines in MRST, it is
sufficient to input only the reservoir density of the fluids.

6.2 Black Oil Simulators


In commercial reservoir simulators, a “Black Oil” model is usually assumed. Up to three
phases may be simulated in this kind of model (oil, water and gas). The composition
of the oil and the gas is not specified (hence the term “black oil”), and the phases are
treated as components. However, gas may be dissolved in oil and water. A typical
table of PVT data for a “dead oil” – i.e. no dissolved gas coming out in the reservoir
(pressure above the bubble point) is given in Table 3. The formation volume factor, Bo,
and the viscosity vary with pressure.

P (psi) Bo (rb/stb) Visc (cP)


1000 1.202 1.15
2000 1.180 1.35
3000 1.167 1.55
4000 1.158 1.75
5000 1.151 1.95
6000 1.145 2.15

Table 3
Example of a typical dead oil table.

Table 4 shows a table for live oil, and includes a column for Rso solution gas-oil ratio.
As pressure increases, more gas can be dissolved in the oil, so Rso increases, and Bo
increases too. Beyond the bubble point pressure, no more gas is dissolved, so Rso stays
constant, and Bo decreases (Figure 33). The oil is under-saturated here in this pressure
range.

©HERIOT-WATT UNIVERSITY December 2017 v2


43

Rs (Mscf/psi) P (psi) Bo (rb/stb) Visc (cP)


0.165 400 1.0123 1.17
0.335 800 1.0255 1.14
0.500 1200 1.0380 1.11
0.665 1600 1.0510 1.08
0.828 2000 1.0630 1.06
0.985 2400 1.0750 1.03
1.130 2800 1.0870 1.00
1.270 3200 1.0985 0.98
1.390 3600 1.1100 0.95
1.150 4000 1.1200 0.94
1.600 4400 1.1300 0.92
1.600 4800 1.1255 0.92
1.600 5200 1.1210 0.92
1.600 5600 1.1165 0.92

Table 4
Example of a typical live oil table (Schlumberger, 2015).

Figure 33
Graph of Rso and Bo vs pressure.

6.3 Compositional Simulators


Sometimes, it is important to know the composition of the oil and the gas phases, and
how the components are interacting during a simulation. For example in a gas flood

©HERIOT-WATT UNIVERSITY December 2017 v2


44

(e.g. CO2 EOR), the gas may mix with the lighter oil components first, then slightly
heavier components, and possibly the heaviest components may be left behind
(asphaltene deposits). If it is necessary to simulate this complex behaviour a full
equation of state (EOS) must be input into the simulator. There are several common
EOS, e.g. Redlich-Kwong and Peng Robinson. These both require several parameters,
which must be obtained from laboratory data. We discuss compositional simulation
further in Topic 9.

7. References
Anderson, W.G., 1986, 1987. Wettability literature survey–part 1 to part 6. J. Pet.
Technology,.1125-1144.

Brooks, R.H. and Corey, A.T., 1964. “Hydraulic properties of porous media and their
relation to drainage design”, Transactions of the ASAE, 7(1), 26-0028.

Corey, A.T., 1954. “The interrelation between gas and oil relative permeabilities”,
Producers Monthly, 19 (1), 38-41.

Craig, F.F., 1971. “The reservoir engineering aspects of waterflooding” (Vol. 3), Society
of Petroleum Engineers.

Jiang, Z., Couples, G., van Dijke, M.I.J., Sorbie, K.S. and Ma, J., 2012. “Stochastic Pore
Network Generation from 3D Rock Images”, Transport in Porous Media, 94, 571-593.

Johnson, E.F., Bossler, D.P. and Naumann, V.O., 1959. “Calculation of Relative
Permeability from Displacement Experiments", Trans AIME, 216, 370-372.

Lie, K-A, 2016. “An Introduction to Reservoir Simulation Using MATLAB: User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, www.sintef.no/mrst.

Nutting, P.G., 1934. “Some physical and chemical properties of reservoir rocks bearing
on the accumulation and discharge of oil”, Problems of Petroleum Geology, 825-832.

Schlumberger, 2015. Eclipse Reference Manual.

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 8 – History Matching


Contents
1. Introduction ............................................................................................................ 3
1.1 What is History Matching? ................................................................................... 3
1.2 Example ................................................................................................................ 3
2. History Matching Overview ........................................................................................ 5
2.1 Data for Modelling................................................................................................ 6
2.2 Sensitivity Studies ................................................................................................. 9
2.3 Choice of Simulator ............................................................................................ 10
2.4 Data to Match ..................................................................................................... 11
2.5 Misfit Equation ................................................................................................... 12
3. History-Matching Workflow ................................................................................. 13
3.1 Material Balance ................................................................................................. 14
3.2 Pressure Matching .............................................................................................. 16
3.3 Water cut or “Saturation” Match ....................................................................... 18
3.4 Matching BHP or THP ......................................................................................... 20
3.5 The Stratigraphic Method .................................................................................. 21
4. Assisted/Automated History Matching Methods ................................................. 22
5. Summary............................................................................................................... 23
6. References ............................................................................................................ 24

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 Explain what is meant by history matching.

 Describe the types of data which are used for history matching.

 Explain the types of parameters which are adjusted during a history match.

 Outline how sensitivity studies may be performed.

 Define the misfit in history-matching.

 Outline a history matching workflow.

 Compare gradient and stochastic methods for history matching.

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction

1.1 What is History Matching?


There is always a lot of uncertainty in a reservoir model. Even if properties are well
characterised around a well, the nature of the reservoir between wells is poorly known.
Therefore, once a reservoir has been on production for a period of time, engineers may
well find that their predictions from reservoir simulation do not match up with the
observed recovery – i.e. the production history. The model therefore has to be
adjusted so that the simulation results agree with the observations. This is known as
history-matching.

Most of the information collected prior to developing a reservoir is static data –


seismic, well-logs, etc. Well test analysis gives some dynamic information, but may be
limited. A lot more information on a reservoir is obtained from dynamic data once the
reservoir has started production – such as the variation in pressure at the wells, and
the production rate of different fluids. For example, if we find that the pressure has
reduced significantly after a few months, this may indicate that there is no aquifer
support and water-flooding is required to maintain pressure. Or if water-flooding is
being used, a pressure drop at a production well could mean that there is limited
communication between this well and an injection well. This could be due to
compartmentalisation of the reservoir due to faulting. Case Study 2 in Topic 1 (Section
4.2), on the Schiehallion Field is an example of this.

The main aim of history matching is to improve the reservoir model so that we can
make better predictions. Realising how a model needs to be altered to match the
history gives a better understanding of the reservoir, and allows engineers to decide
how to manage the reservoir in the future. (See Topic 1, Section 3.2.) There are some
problems, though. History matching is an inverse process – we need to determine
what model can produce the observed restuls. Also, it is non-unique. This means that
we could alter the model in different ways, and obtain a match each time. We don’t
know which is correct until further production data is obtained. Also there are so many
uncertainties, we need to decide which should be adjusted to obtain a sensible match.

Engineers spend a lot of time on history matching. But, using the workflow presented
in this Topic should help to achieve a match within a reasonable time frame.

1.2 Example
Figure 1 shows an example of how history matching is carried out.
1. We start with an initial model, in which some of the parameters may be rough
estimates (e.g. the average permeability of each layer of a reservoir model).
2. We run a simulation (sometimes this is referred to as a “forward model”).

©HERIOT-WATT UNIVERSITY December 2017 v2


4

3. We compare the results (production rates and/or pressures) with the observed
production data.
4. We adjust the model (update parameters).

Steps 2 – 4 are repeated until we have an adequate match.

Figure 1
The history matching loop, and some examples.

The graphs on the right side of Figure 1 show examples of history matched data. Water
production rate and Gas Oil Ratio have been measured (blue dots) and are compared
with simulation results (black lines). The initial model was probably wrong, with water
breaking through too early or too late, GOR may have been too high because the
pressure draw down was too great (or vice versa) and so on. The engineer changed the
properties in the model that were considered to be linked to the errors and then
another simulation was run.

In a nice easy world this would happen in one iteration of the loop but due to the
reservoir complexities it is most likely that the first guess will be wrong and many
simulations will be required. Some models will be worse than before and the engineer
can become quite frustrated. Eventually though, with application of some sound
principles (see below) the headaches can be avoided and we can get the kind of match
seen in the data shown.

History matching is not the final aim, however. Remember that we want the model to
predict the future as well as the past so that we can explore various decisions to be
made about the reservoir.

©HERIOT-WATT UNIVERSITY December 2017 v2


5

We may collect a set of models that match the data. In the Figure 2 there is a family of
models that give a good match. Models that don’t match very well have been ignored
here. We then run the models forward in time to examine their behaviour. We note
that uncertainty remains and there is a spread in the forecast. There are several
reasons why the models are still uncertain. Firstly the history matched models are only
being affected by the properties of the reservoir that influence the regions producing
during the history period. Well pressures are only affected by regions in pressure
communication with the wells where we have data. Water production data reflects the
regions where the water has passed through and is in communication with. Secondly,
the different processes depend on different properties. For example depletion
depends on the overall connectivity of the reservoir. A waterflood depends on
variations in connectivity. A gas flood depends on vertical AND horizontal permeability.
So a good match to history does not mean that the forecast will be accurate. In the
above example, a range of outcomes exist.

Figure 2
Example comparison of results that match history, and subsequent forecasts.

2. History Matching Overview


Fundamentally history matching can be considered as a set of numerical experiments
where we test various hypotheses. We can explore whether or not there is aquifer
support, what the flow paths are in the reservoir, and where are we likely to leave oil
behind. We need to get all of these correct to obtain a good history match. We
therefore gain a better understanding of the reservoir processes. We also find out
whether or not we have sufficient data. Do we need more relative permeabilities or
are there processes that we have not properly represented in the model for lack of

©HERIOT-WATT UNIVERSITY December 2017 v2


6

data? We may identify that there are some unusual operating conditions, particularly
near the wells. We certainly get a better understanding of what is likely to be
happening and we can explore this in an objective manner. Ultimately of course we
want to be able to forecast the future behaviour with improved confidence.

Before we carry out history matching we must answer some fundamental questions.
The first question is really what do you want to use the model for once you get a
match? Do you want to simulate the continued behaviour of the field as it is or do you
want to change the nature of the field by drilling more wells, using an EOR process or
otherwise changing the facilities? These decisions alter the nature of the flow
behaviour and we need to make sure that the model we get from history matching will
be capable of also being used if the conditions change.

We need to decide what it is we want to simulate because that influences the kind of
model we will use. Early in a field life there is a lack of data so there is no point in
constructing and running very complex detailed models. While these may be precise
in their representation of the reservoir, they will take too long to run and will inevitably
require more runs as well to figure out the details. If we don’t need particular details
we leave them out.

We also need to decide what it is that we want to change in the model. This ties back
to the first questions but it also depends on the data we are working with. The
geologists, petrophysicists and geophysicists have a part to play in helping answer this.
They can also give us an idea by how much the various properties should be varied.

We should have a clear definition of how we will define a ‘good match’. We will see
that we can do this very objectively or we can eyeball it from graphs. We should also
decide which data we trust. Which data has been measured accurately and are there
effects in the data that render it less reliable?

2.1 Data for Modelling


First we need to consider the data that we have used to generate the model. Some of
it may be considered as “hard” data. A measurement of permeability in a piece of rock
is hard data. We ought to be pretty certain about that, along with its porosity. Log data
are also hard data (e.g. Figure 3, top). We measure relative permeability data and
capillary pressure curves (e.g. Figure 3, bottom), properties of the rock and the fluids,
and we may also have measured where the fluid contacts are. If we know what they
are we use this information and do not change it.

Of course there are issues of representivity. The rock we have measured no longer
exists in the reservoir and we produce from other bits of rock. Is it the same? There

©HERIOT-WATT UNIVERSITY December 2017 v2


7

are also scales of measurement that we should concern ourselves with. A simulation
grid is probably 25 to 100 m long and 0.5 to 5m thick. The sample rock is measured in
centimetres so the effects of scale can be important.

Figure 3
Top: Example of log data. Bottom: Example of relative permeability data.

We also use soft data. Seismic data (Figure 4, top left) should be considered to be soft
data. It lacks resolution for one thing, but there can be uncertainties in its calibration
which mean it may be good at giving a relative measure but absolute values are less
reliable. Porosity and permeability relationships (Figure 4, top right) are often noisy,
which introduces uncertainties. The geostatistical model that the geologist creates also
invokes soft relationships (Figure 4, bottom). Geostatistical information such as semi-
variograms are often obtained from analogue outcrops or fields. They give a good
representation of the spatial variation in the reservoir but they are not precise.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

Figure 4
Examples of soft data. Top left: seismic data. Top right: permeability-porosity cross-
plot. Bottom: conceptual geological model.

Then we have the complete unknowns. The petrophysical properties in grid cells 300m
from a well could vary considerably from the measured values at the wells. We
therefore have to treat many grid cell properties as unknowns (although there are
limits to the uncertainty). We define these as the known unknowns such that these are
properties that we know exist but we just don’t know what they are. We can also miss
certain properties such as sub-seismic faults or shale layers. We don’t know they are
there and we certainly don’t know their properties.

In summary, we need to honour the hard data, be guided by the soft data, and alter
some of the parameters which are unknowns, as described in the next section.

©HERIOT-WATT UNIVERSITY December 2017 v2


9

2.2 Sensitivity Studies


Before we begin history matching though we need to identify what we need to change.
As mentioned previously, the geoscientists should be able to give ideas on what
uncertainties exist in the data that have been used to construct the model. This is
useful because we don’t need to change highly certain data. This is not all that we need
however. Not every uncertain parameter will have a significant influence on the
production behaviour so it is important to perform a sensitivity analysis to determine
what is important. Usually this means choosing a base case simulation and changing
one parameter at a time and looking at the response. In the spider diagram in Figure 5
(top), three parameters are considered and multiple parameter values are selected
though often an upper and lower value of each is all that is needed. In the case shown
here, kv:kh is least important and could be ignored in a history matching study (at least
to begin with). In Figure 5 (bottom, taken from a study of the Schiehallion Field), a
number of different properties are considered. An upper value and a lower value were
chosen and the simulator is run for each. This enables the response to be calculated
relative to the base case. Once the variations have been calculated from the simulator
the bar chart can be plotted and the largest bars indicate the dominant parameters.
Note that this figure is often ordered so that the smallest bars are at the bottom and
creates a tornado appearance so this is called a “tornado chart”.

Once the important parameters are identified we can start to run models by changing
these in different combinations.

©HERIOT-WATT UNIVERSITY December 2017 v2


10

Figure 5
Top: Example of a spider diagram. Bottom: Example of a sensitivity plot as a bar
chart.

2.3 Choice of Simulator


It is also important to choose the correct simulator, or you could just use material
balance (see Section 3). For history matching of a young field (a few years old), a
simpler model will suffice to minimize run time and the number of unknowns. Further,
if only water flooding has been carried out and there is little gas evolution, a streamline
simulator (Topic 6) could be used which can speed up the simulation by up to a factor
of 10 (Figure 6). Ultimately though, for older fields at least a Black Oil simulator is
needed and if EOR has been carried out a compositional simulator (Topic 9). Choosing
the right balance between precision and speed is important.

Figure 6
Example of a streamline simulation.

©HERIOT-WATT UNIVERSITY December 2017 v2


11

2.4 Data to Match


An important question for the engineer is “what data do we match on?”

Before we answer that, we should also consider that the simulation must be set up
with a set of well controls. That is we must choose to define either rate or pressure
controls for all wells along with equivalent pressure or rate limits. The history matched
model requires this the same as any other. It obviously makes sense to use the history
data in this context. Thus, in the exception of extreme problems with the model, the
control data will always be perfectly matched. For example, we might instruct the
reservoir to produce from wells at the historic liquid oil rates with some defined
pressure minimum (in some fields this might be the bubble point if it is known that that
has been avoided). Then should the simulation reach that pressure minimum, the
result of the model is obviously very bad and that version of the model should be
avoided.

So what data should we control on and what should we match on? Logically, the most
accurate data should be used as the control. This is usually the field production volume
of oil over each simulated time interval. This is what we sell after all. Individual oil rates
may be specified for each well but these are often obtained through the process of
allocation and may not be accurate.

Traditionally history matching was carried out manually. Figure 7 shows a ‘good match’
from a North Sea reservoir as determined by the operator. The engineer controlled the
simulation using Liquid Rate or Oil Rate and then compared pressures and water cut
(or water rate).

The water cut was the better matched of the two. The breakthrough time was slightly
late but the build-up in water cut was good despite a discrepancy around 350 days and
a divergence at 600 days. The pressure match was still acceptable if not quite as good.
The early pressures were matched quite well as well as the later cluster of pressures at
around 600 days. There are clusters of points (around 300 to 350 days and 400 to 450)
where the trend was wrong.

©HERIOT-WATT UNIVERSITY December 2017 v2


12

Figure 7
Comparison of history and best-matched model for bottom-hole pressure (left) and
water cut (right).

Given Figure 7, the engineer has had to decide which of the points in the curves were
good and which should be ignored. The production engineer would have been asked
for any known problems such as sand production, changes to the well controls that
were important locally. The data uncertainty (i.e. how accurately it was measured)
should have been considered also. Could the engineer have drawn error bars on the
graph, for example? Do we have enough information to do that? These all help decide
the accuracy of the match.

Could we try to improve the match further? If we modify the model to improve some
of those discrepancies mentioned above, we may find that while we improve, say, the
water cut, the pressure actually gets worse (or vice versa). It is quite possible that this
is what the engineers observed and this is why they stopped at this point. When we
get to this point we have reached the limits to which the model can be improved. We
require additional detail to improve it. Perhaps the model is too coarse or the
geological conceptual model is incorrect. Pragmatism may force us to stop at this point.

2.5 Misfit Equation


It is usually the case that the data types have different levels of measurement precision
AND we may make one match better at the expense of the other. An objective
measure of match to the data is therefore required. This measure is also used in
‘automatic’ methods of history matching that we will discuss later.

The difference between the model results and the measured data is calculated for
every comparable time-step, variable and well that we want to include here. This

©HERIOT-WATT UNIVERSITY December 2017 v2


13

creates what we might think of as a residual vector. If we calculate the magnitude of


that vector we would sum the square of the residuals and we could use that as a
quantified measure of misfit. The one problem with that is that pressures are
measured in 100’s of PSI while water cut may be measured in fractions. The misfit of
the pressure will be much larger based on the units of pressure alone. We therefore
need to scale the data somehow.
A sensible approach is to scale according to the magnitude of the data errors. Thus if
we know the standard error of measurement for a variable, we can use that. This is the
 in Equation (1). By weighting according to the data errors in this way, we match the
more accurate data more closely. There is also a statistical basis for this found in
Bayesian theory but we won’t go into that for now. This leads to a misfit equation like:

X  X ijksim 
obs 2
N w N var Nt
Misfit    
ijk
(1)
i 1 j 1 k 1  ijk2

"X" is a variable that we both observe and predict (indicated by superscripts "obs" and
"sim"). There are Nw wells, Nvar variables and Nt time steps and these are counted by i,
j and k respectively. ijk2 is the data error. Usually this is assumed to be a constant for
each variable across the various time steps and wells but we include the "i" and "k"
indicators for generality.

In some cases we don’t have an accurate measure of data accuracy so the misfits for
individual variables are simply weighted arbitrarily, or at least estimated. This is still
useful but we have to be careful about the resulting model and should not just accept
the best one only because the measure of the best model is itself uncertain.

3. History-Matching Workflow
The main idea is that we try to reduce the complexity of the problem and work with an
organising principle. It would be a nightmare to try and change everything at once so
we try to identify parameters that control certain parameters first.

History matching follows the following steps:

 Material balance: We start by trying to ensure that we have the right average
properties to get a good material balance in the model. We use the material
balance equation in place of the simulator at this stage but derive properties to
constrain the simulation.

 Match the pressures (field/reservoir): We use the simulator to try to get the
energy of the system right and match the field average pressures.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

 Match the GOR/water cut: With those matched we can move on to more local
properties and match the water cut/rate and/or the GOR.

 Match the BHP or THP (PI): Local variations around the wells are made to match
Bottom Hole or Tubing Head Pressures.

Each of these steps is discussed in detail below.

3.1 Material Balance


Material Balance is discussed in detail in Topic 2, Section 2. For further details see Dake
(1978 or 1994) or Ahmad (2006) or any good reservoir engineering text book. The
material balance equation is well understood and has been used by engineers for
decades. In essence the material balance equation is just a simulation with its own
parameters, assumptions and approximations. We use it to estimate the oil in place
but we can also determine the aquifer influx and other drive mechanisms. The real
advantage of it is that it can be set up in a spreadsheet in some cases. There are also
material balance software tools that are used in more complex compartmentalised
reservoirs.

Material balance works on the simple observation that:

volume produced = volume expansion + volume influx.

As discussed in Topic 2, the Material Balance equation requires data. You need to have
history data telling you how much oil and other fluids have been produced as well as a
measure of the average field pressure from shut in measurements or by estimation
from flowing pressures. The correct PVT data is also needed to relate changes at
various pressures. Note that the material balance only applies to those volumes in
pressure communication with the wells. That may include part or all of the aquifer
depending on how you model it. The material balance equation is comparable to a
geological model, it just averages everything over the flowing volume. If you find there
are differences between what the geological model predicts and the equation, the
model is probably not representing internal barriers. The geologist and the engineer
have to work it out.

In Topic 2 we introduced the material balance equation for primary depletion where
we stay above the bubble point and there is no influx from an external aquifer. (There
may be a “water-leg” though and this can contribute to the expansion of water).

©HERIOT-WATT UNIVERSITY December 2017 v2


15

N  Boi Boi  Swi  cw  c f 


1  p     p (2)
 N  Bo Bo  1  S wi 

The symbols have the usual meanings. (Check Topic 2, if you are uncertain.) Note that
it is implicitly time dependent, oil expansion is included in the oil formation volume
factor, the average saturation refers to all water in pressure communication with the
wells and p is the increase in pressure which is negative under depletion!

Figure 8
Plot of the right side of Equation (2) over the drop in pressure, -P.

In the first tutorial for Topic 2, we used this equation in these conditions to estimate
the original oil in place. In other primary depletion scenarios, either with water influx
from an aquifer OR gas evolution, additional terms are required in the equation.
Additional compartments in the reservoir may be in weak communication also
requiring separate equations for each one and a transfer term to allow flow between
them.

Equation (2) is used to calculate STOIIP, N by evaluating the right hand side for a given
pressure drop and then solving for N.

As it stands Equation (2) can be used to solve the material balance for a reservoir in
communication with an aquifer provided that the correct average saturation is used to
include the aquifer volume. In general though, a large aquifer will not reach pressure
equilibration fast enough and so additional influx terms are required. These are also
needed for injector wells.

The mass balance equation can be written in a more general form to include water and
aquifer influx as well as oil and water expansion (extendable to include gas cap
expansion):

©HERIOT-WATT UNIVERSITY December 2017 v2


16

F Bw We  Winj 
N (3)
Eo  E f , w Eo  E f ,w

where F is the volume withdrawn in rb, N is the STOIIP, We is the water influx volume
from the aquifer (here in stb, but was in rb in Topic 2), Winj is the known water injection
volume, Eo is the oil expansion and Ef,w is the formation and water expansion term
obtained from the PVT data as in the RHS of Equation (2). (Both Eo and Ef,w are in units
of rb/stb.) (See, for example, Dake, 1994). This equation basically says:

volume withdrawn influx


=STOIIP+ .
expansion expansion

The left side of Equation (3) is then evaluated and plotted over time. The response
depends on the strength of the influx and can be used to infer influx rates (Figure 9).
A flat line indicates zero influx and we can calculate the oil in place. We can also
estimate the influx volumes from the shape of the curve and use these to constrain the
simulator later. A weak aquifer follows the red curve while the blue curve indicates a
good aquifer.

Figure 9
Schematic graph of the right side of Equation (3) over time.

3.2 Pressure Matching


Once the material balance has been carried out we can then start using the full field
simulation model to match field average pressures. While the simulation is more
complex than the material balance equation and we cannot invert it directly, we are
using it in a very similar manner. We seek to calibrate the overall volumes of the
reservoir, the energy and the drive mechanisms. Of course the simulation model has
spatial distribution of properties and flow behaviour and our main aim is to determine
these more accurately than the averages obtained from material balance.

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Figure 10
Example of the average pressure decline in a reservoir, comparing simulation with
history.

The average pressure depends on the material balance of the system. Pressure
declines if we produce more than we inject (e.g. Figure 10). The rate of decline is
governed by a number of variables including how much material exists that can expand
to replace produced volumes and the rate at which it expands (i.e. compressibility).

Gas, oil, water and the rock can all expand as the pore pressure reduces. We therefore
need to know how much of each is in pressure communication with the wells. The
amount of oil depends on the available pore volume, the depth of the contacts, the
end points of the relative permeability curves and also the strength of any capillary
transition zones (Topic 7, Section 4.1.3).

The volume of water includes connate/irreducible water, which can expand, water in
the transition zone and any water below the oil water contact (OWC) that is in pressure
communication with the well. The volume of gas depends critically on the depth of the
gas cap as well as the total pore volume which contains the gas. The location of faults
and their transmissibility also affects the pressure decline. If sealing, they define the
available volumes. Non-sealing faults also act to dampen expansion in poorly
connected regions.

In the schematic example in Figure 10, the observed field average pressure declines
quickly initially but then starts to flatten out. If there is no pressure support at all, the
field enters pseudo-steady state and the pressure declines at a constant rate such that
offtake is matched by expansion. The pressure decline would be a straight line. An
aquifer helps increase the volume available for expansion and so the rate of decline of
pressure slows down once it starts to flow. An infinite aquifer would result in a flat
pressure plot eventually. A moderate aquifer seems to be providing some support in
the example. The simulation does not see such a rapid decline, however, and so there

©HERIOT-WATT UNIVERSITY December 2017 v2


18

is too much energy in the modelled system or too much fluid connected to the well.
Some of the above model parameters need to be altered. We determine which by trial
and error.

There are some key steps to take to getting a reasonable field pressure match. Firstly
the offtake of the reservoir should be correct. We usually control the wells using the
liquid rate at this stage and we may even choose to set the rate in reservoir units (rb
instead of stb). If we use the surface units and the pressures are wrong at the well
then the formation volume factor may give an incorrect reservoir volume offtake.

The average pressure from the reservoir is then compared to the model. We take care
not to alter the properties obtained from the material balance calculation but we
change the pore volume and aquifer properties. The main issue here is likely to be the
spatial distribution of the pore volumes and in particular the transmissibility in the
reservoir which can control the rate of flow.

Note that this step is not the same as matching on BHP which is a flowing pressure.

3.3 Water cut or “Saturation” Match


The water cut and/or GOR match depend essentially on two things:

1. How close is the source of water or gas to the production well(s)?


2. How good is the sweep and displacement in the model compared to reality?

The source of the invading fluid depends on the location of the well relative to fluid
contacts as well as the location of any barriers.

Other common properties to change at this stage are the permeability and relative
permeability data. Permeability variations tend to alter the sweep (i.e. where the flood
front reaches). On the other hand the relative permeability affects displacement (i.e.
how much of a grid cell gets swept out relative to the amount of fluid that invades).
(Think about Buckley-Leverett displacement, here.)

It may be tempting to use different relative permeability curves in different regions.


This needs to be justified. Do you have other information that supports this approach
(e.g. geological information)? Otherwise keep it simple and use a small number of
relative permeability curves.

Permeability variations should also be justified geologically. Is the geologist happy if


you place a channel or a fault in the reservoir to get the match for example?

©HERIOT-WATT UNIVERSITY December 2017 v2


19

It is also good to try to break down how you change the model parameters based on
your understanding of flow behaviour. For example if you suspect coning then vertical
permeability is likely to be the first place to consider making changes. You may have
to alter or place shale layers and the well relative permeability curves may be
important here. For lateral sweep the horizontal permeability and layer thickness can
be very important (but don’t forget vertical permeability as it controls gravity slumping
and crossflow between layers). Pore volume is often important in this case as a more
porous rock takes longer to sweep. The relative permeability data have a strong
influence on both breakthrough and, through displacement efficiency, the build-up of
water production.

Figure 11
Schematic graph of water cut measurements and predictions showing two possible
miss-matches.

In Figure 11 (left), the simulation predicts early breakthrough and faster increase of
water cut. We need to consider the nature of the flow to help guide us to decide on
what we should change. In the case of horizontal flow we may have placed a thief zone
(high permeability layer) in the model where we should not have. We may also require
a shale barrier of some sort (depending on geological interpretation). In the case of
vertical flow we might need a may require a shale layer to be added to slow down the
movement of the front.

The sort of mismatch represented in Figure 11 (left) also depends on the shape of the
relative permeability curve. If the water relative permeability is too high we would see
earlier breakthrough (see Buckley-Leverett analysis) and greater production. We may
have overestimated the degree of oil wettability, but it depends on how the relative
permeability data was obtained as well.

The early breakthrough is also characteristic of too much dispersion in the model. This
can be a numerical effect if the front is spreading because the cells are too large or the
flow is aligned to the grid. Either pseudo relative permeabilities may be required or a
refined grid. The model could have too much physical dispersion, though that is
equivalent to the thief zone mentioned above.

©HERIOT-WATT UNIVERSITY December 2017 v2


20

Note that the above curves will probably cross eventually. The key here is to be able
to break down what may be affecting flow so that you don’t consider everything to be
important.

In Figure 11 (right) the breakthrough time is correct but the later water production is
wrong. Be careful here because the correct breakthrough could be the result of two
wrongs making a right. For example the wrong physical dispersion mixed with the
wrong permeability contrast in layers may have accidentally resulted in the right
breakthrough time.

On the other hand the model may be predicting the correct location of the flood front
but the displacement is wrong. This is usually because the ratio of relative permeability
is wrong. Another possibility is that we have located the aquifer correctly relative to
the wells but it supplies too much water. We would of course expect to see some effect
of that in the pressure data but we may have compensated for that with an error
elsewhere.

As you can see then there are many possibilities that affect how the fluids move
through the model of the reservoir and it is hard to get them right based on production
data alone.

3.4 Matching BHP or THP


This is the last step of history matching and we consider changing properties very local
to the wells. We can leave the BHP to the last provided that we are not affected by
pressure dependent variables during the other stages.

We have to go through this step before we get into forecasting (predicting beyond the
history period). This is usually because well controls are set to pressure limits during
forecasting rather than production rates.

The properties that are important are usually those that affect the productivity index.
The local permeability, the completion length or the effective radius (See well
modelling in Topic 3, Section 4) all affect flow here. We may also consider relative
permeability changes equivalent to the application of pseudos, depending on
heterogeneity effects around the wells (Topic 6, Section 5.3).

Figure 12 illustrates the complexity of this stage. We see a cross section through a
model that includes some deviated wells. These wells pass through the cells. A
modelling package like Petrel can be used to estimate the true length of completions
for each of the cells that the wells pass through. This is then used to estimate the

©HERIOT-WATT UNIVERSITY December 2017 v2


21

productivity index (PI). The measured completion data and the location of the well may
be incorrect. There may be formation damage or even well collapse for example that
alters the PI. The PI also depends on the individual permeability of each cell, itself an
estimate through upscaling of core/well measurements and lateral interpolation.

Figure 12
Cross-section of a model indicating well trajectories and completions (green
symbols).

3.5 The Stratigraphic Method


On the whole, history matching is quite a complex process. The key is to avoid being
overwhelmed by the enormity of the problem. Potentially the transmissibilities, pore
volumes, relative permeabilities of every grid cell may be wrong and require changes.
It helps therefore to organise our understanding to reduce the number of unknowns
(degrees of freedom) in any way we can.

The Stratigraphic Method (Williams et al, 1998 and Ertekin et al., 2001) is one such
method (illustrated in Figure 13). The aim is to focus on global properties first (average
permeability, porosity, total volumes etc. as well as the boundaries). Focus on those
and get a match that is as good as possible. Then break the problem down to a smaller
level. We may consider “flow units” which is another way of identifying regions in the
model that are similar. Often geological information helps here. We may identify
different facies or groups of layers considered to be similar. Individual layers can then
be considered and finally sub-volumes around the wells. This process takes a lot of pain
out of a less organised and often random approach.

©HERIOT-WATT UNIVERSITY December 2017 v2


22

Figure 13
The Stratigraphic Method (Williams et al, 1998).

4. Assisted/Automated History Matching Methods


Manual history matching is quite time consuming. With faster and cheaper computers
we can now run many simulations and use optimisation algorithms to select new
parameters for us. Assisted history matching is a broad term used to identify more
mathematical and statistical methods of assessing reservoir behaviour during history
matching. There has also been a growth in the number of software packages that can
be used for this.

The first step in history matching is usually to understand what controls flow and the
history match, so sensitivity studies are required to determine the most influential
parameters.

Once we’ve eliminated some properties we can then carry out Automated History
matching where an optimization algorithm tries to find the best set of parameters for
the model so we get the best match. The key here is to remember that automatic
history matching helps us identify by how much we should change reservoir
parameters but it doesn’t tell us what to change. The sensitivity analysis helps with
that decision. We also have to query the results quite carefully.

Automatic history matching methods can then be split into several approaches.

Gradient methods are deterministic. Based on a starting model there is one outcome,
hopefully an improved model. The approach is usually fairly fast. A typical example is
the method used in Schlumberger’s SIMOPT. Application of the method does require
calculation of gradients (i.e. how much the misfit changes with a change in a
parameter).

©HERIOT-WATT UNIVERSITY December 2017 v2


23

Gradient methods use the same approach as used within a reservoir simulator to
improve the search for a numerical solution (See Newton-Raphson in Topic 5, Section
6). Starting with a model, the rate of change of misfit is calculated relative to the
change of each parameter. This step is the difficult part. Based on these gradients and
the misfit itself, a new model is estimated using a Quasi-Newton approach (the misfit
is assumed to be linearly dependent on the parameters). Once the new model has
been run and a new set of gradients and the misfit have been calculated the process is
repeated. This goes on until the process converges on a result.

The advantages of the approach are that the outcome is always the same and it is quite
fast. The disadvantage is that there is only one outcome and the choice of starting
model may influence that. There may be several optimal solutions and we might not
find the best one. Gradient methods were very popular early on with automated
history matching but have become less popular.

Multiple model stochastic approaches require many more models. They are more
costly but are more likely to find the best outcome. In fact the advantage is that they
may find many good outcomes and also enable an analysis of the behaviour.

Multiple model stochastic methods are more common. A great advantage is that they
are easier to set up for any given problem and the results are more satisfactory even if
they are more computationally intensive. Once the parameterisation has been
defined, a number of models are generated by randomly picking parameter values. All
models are simulated and misfits calculated.

The next step is to create a new set of models using the information we already have
so that the new models are more likely to have lower misfit. This step is different in
different approaches. Models such as evolutionary and genetic algorithms work by
ranking the models, picking the best ones obtained so far. These are then mixed
together in the hope that the new combination provides a better model and the
process is repeated. Gradually we get to the stage were no more models are generated
with a better misfit and we stop. A further advantage of this approach is that we get
a lot of useful information from the statistics of all the models as well.

5. Summary
To summarise: changes to the model should be kept simple. Do not invent properties
of behaviours that are unphysical or “ungeological”. Use your common sense.
 Start with material balance and assess the reservoir behaviour. From that, get the
global representations of the drive and the volumes right.

©HERIOT-WATT UNIVERSITY December 2017 v2


24

 Then do the field (average) pressure match. This is a repeat of the material balance
analysis, except the model is more complex. At this stage you can adjust the pore
volume distribution within the reservoir, transmissibilities and aquifer properties.
 Then match watercuts and GOR. Parameters which can be varied at this stage are:
permeabilities, transmissibility barriers (e.g. due to faulting), kv/kh ratio and
relative permeabilities
 Finally match the BHP or THP with local changes at the wells, adjusting factors
which influence the well PI, and near well relative permeabilities.

Throughout this procedure, the geologist and engineer should work together to ensure
that the best possible results.

6. References
Ahmad, T., 2006. “Reservoir Engineering Handbook”., Third Edition, Elsevier.

Dake, L.P., 1978. “The Fundamentals of Reservoir Engineering”, Developments in


Petroleum Science, 8, Elsevier.

Dake, L.P., 1994. “The Practice of Reservoir Engineering”, Developments in Petroleum


Science, 36, Elsevier.

Ertekin T., Abou-Kassem J.H., King, G.R., 2001. “Basic Applied Reservoir Simulation”,
SPE Textbook Series, Volume 7.

Williams, M.A., Keating, J.F. and Barghouty, M.F., 1998. “The Stratigraphic Method: A
Structured Approach to History-Matching Complex Simulation Models”, SPE 38014,
SPE Reservoir Evaluation and Engineering, 1 (2), 169-176.

©HERIOT-WATT UNIVERSITY December 2017 v2


1

Topic 9 – Simulation of Reservoir


Recovery Processes
Contents
1. Introduction ................................................................................................................ 3
1.1 Note on EOR ......................................................................................................... 3
2. Compositional Simulation .......................................................................................... 4
2.1 Introduction .......................................................................................................... 4
2.1.1 Masses of Oil, Water and Gas in a BOS Model .............................................. 5
2.1.2 Mass of a Component in a Compositional Simulation .................................. 6
2.2 Gas Injection ......................................................................................................... 8
2.3 CO2 Storage......................................................................................................... 11
3. Chemical EOR............................................................................................................ 13
3.1 Polymer Floods ................................................................................................... 14
3.1.1 Buckley-Leverett Theory and Polymers ....................................................... 14
3.1.2 Numerical Simulation Example.................................................................... 16
4. Thermal EOR ............................................................................................................. 17
5. Simulation in Fractured Reservoirs .......................................................................... 18
5.1 Introduction ........................................................................................................ 18
5.2 Dual Porosity Method......................................................................................... 19
5.3 Discrete Fracture Models ................................................................................... 21
6. Closing Remarks........................................................................................................ 24
7. References ................................................................................................................ 25

©HERIOT-WATT UNIVERSITY December 2017 v2


2

Learning Objectives
After studying this topic, you should be able to:

 Explain the difference between black oil (BOS) and compositional simulation (CS)
o Calculate the mass of components in BOS
o Calculate the mass of a component in CS
o List examples of processes which can be simulated with BOS or CS

 Describe the processes which arise when CO2 is injected into a reservoir or
aquifer

 Explain how the use of polymer improves recovery, using Buckley-Leverett theory

 Explain why simulation in fractured reservoirs is important


o Describe the main problems regarding simulation of fractured reservoirs
o Describe the dual porosity model
o Outline methods for explicitly modelling fractures

©HERIOT-WATT UNIVERSITY December 2017 v2


3

1. Introduction
So far in this course, we have dealt with simple processes for oil recovery – draw down
and water flood. We could also use the methods discussed to simulate recovery in a
gas reservoir (either with just drawn down, or water flood). However, there are many
more complex processes involved in reservoir recovery. Engineers also need to be able
to simulate these techniques. In this Topic we consider a range of recovery processes
and how they are simulated. The aim of this topic is to make you aware of these
advanced processes, although you will not be required to run simulations.

In particular, enhanced oil recovery (EOR) is used in many reservoirs. The term EOR is
used when engineers inject materials not normally present in the reservoir to increase
recovery (Lake, 1989). This includes gas injection and polymer injection. EOR methods
can be very costly, and so, before deciding on a particular type of EOR, engineers need
to evaluate the cost effectiveness. Reservoir simulation is naturally an important tool
in this case. In Topic 1, we discussed how reservoir simulation can be used in a mature
field. An example of this may be a simulation study to compare the merits of
continuing with a water flood, drilling more producer wells, or using an EOR method,
such as gas injection, or WAG (water alternating gas).

In addition to discussing the injection of CO2 into reservoirs for EOR purposes, there is
also a section on injection of CO2 into deep saline aquifers for storage purposes, in
order to reduce the levels of CO2 in the atmosphere. This is becoming an important
application of reservoir simulation.

We also consider some special types of reservoirs. Firstly, many reservoirs particularly
in Canada and Venezuela, contain “heavy oil”. This oil is particularly viscous, so
recovery is poor with normal water flooding. One solution is to perform thermal EOR,
which involves injecting steam into a reservoir. Another important type of reservoir is
a fractured reservoir. It is estimated that 60% of the world’s oil reservoirs are in
fractured rocks. In such reservoirs conventional reservoir simulation breaks down
because of the huge difference in scale between the fractures (~ 10 m) and the matrix
blocks (~ m or larger). Also, the permeability of fractures is generally very high (10s of
D) compared with the permeability of the matrix. For this reason, special simulation
approaches have been developed for fractured reservoirs.

1.1 Note on EOR


There are basically three different types of EOR: gas injection, chemical injection and
thermal (e.g. steam injection). There are many varieties within these three categories.
Oil recovery may be improved by one or more of the following methods:
 Improvement of microscopic sweep efficiency by overcoming capillary forces at
the pore scale;

©HERIOT-WATT UNIVERSITY December 2017 v2


4

 Increase in the mobility of oil by reducing the viscosity (due to heat or dissolution);
 Improvement of macroscopic sweep by lowering the mobility ratio, e.g. by
increasing the viscosity of the displacing fluid. This helps to improve recovery in a
heterogeneous reservoir.

The EOR method to use depends on a number of factors, such as type of oil
(light/heavy), reservoir conditions (pressure and temperature), reservoir
heterogeneity, availability of gas/chemicals to be injected, and economics.

Figure 1 shows a useful classification of some EOR methods. Here we will focus on two
methods: gas injection and polymer injection, and also mention a few other methods
briefly. First, though, we consider compositional simulation, which is required for gas
injection when it is miscible or near-miscible.

Figure 1
Different types of EOR methods (Spooner, 2016). (WAG = water alternating gas;
MWAG = miscible WAG, SWAG = simultaneous WAG, ASP – alkali surfactant polymer;
SAG = Surfactant and Gas.)

2. Compositional Simulation

2.1 Introduction
Compositional simulation was briefly mentioned in Topic 7, Section 6.3. In
compositional simulation, we keep track of the individual fluid components, which may
be in the oil, gas or water (aqueous) phases. To appreciate the difference between

©HERIOT-WATT UNIVERSITY December 2017 v2


5

black oil simulation (BOS – which we have been using) and compositional simulation
(CS), we compare the calculation of mass of different phases or components in a grid
block.

2.1.1 Masses of Oil, Water and Gas in a BOS Model


BOS simulation 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 need to be entered as tables of data. See Table
4 in Topic 7.

A simple schematic of a grid block in a black oil simulator is presented in Figure 2,


showing the 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. (In
Topic 4, where we calculated the masses of oil and water to derive the flow equations,
we ignored dissolved gas.)

Figure 2
A grid cell in a black oil simulator showing the amounts of oil, water and gas present.
Vp is the pore volume = block volume x ; osc, wsc and gsc are the densities at
standard conditions (60oF and 14.7 psi); Bo, Bw and Bg are the formation volume
factors; Rso and Rsw are the gas solubilities (or solution gas/oil and gas/water ratios).

The equations in Figure 2 are repeated below:

©HERIOT-WATT UNIVERSITY December 2017 v2


6

V p So osc
Mass of oil = (1)
Bo

V p S w  wsc
Mass of water = (2)
Bw

 S g So Rso Sw Rsw 
Mass of gas = V p  gsc     (3)
B Bo Bw 
 g

(The symbols are given in the caption for Figure 2.)

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-alternating-gas


(WAG).

 Capillary imbibition processes.

2.1.2 Mass of a Component in a Compositional Simulation


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 indicated
in Figure 3 (which should be compared with Figure 2). 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 near-critical fluid 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.).

©HERIOT-WATT UNIVERSITY December 2017 v2


7

Figure 3
Phases and components taken in compositional simulation. Cij is the mass
concentration of component i in phase j (j = oil, water or gas). The dimensions of C
are mass/unit volume of the phase; Vp = pore volume = block volume x .

3
Total mass of component i in all the phases = V p S C
j 1
j ij . (4)

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. (See G19RA, Topic 4,
Section 4.3.)

 Gas recycling processes in condensate reservoirs. (See G19RA, Topic 4, Section


5.2.)

Compositional simulation will obviously take longer than black-oil simulation, because
there are more unknowns, namely the concentrations of the different components in
the different phases, and the pressure in one phase (pressures in other phases can be
calculated from Pc, which will tend to zero as miscibility is achieved). There are actually
100s of components in crude oil, but compositional simulation usually focuses on a few
common components – the lighter components – and the heavier ones are lumped
together into “pseudo components”. Typically compositional simulation considers
about 7 components and pseudo components.

©HERIOT-WATT UNIVERSITY December 2017 v2


8

2.2 Gas Injection


Different types gases may be used for gas injection, depending on what type of gas is
available and therefore what is most economical – e.g. methane, nitrogen, air, or CO2.
In Sections 2.1.1 and 2.1.2 we mentioned “immiscible” and “miscible” gas injection.
Oil and gas are immiscible at low pressure and miscible at high pressure. The actual
pressure depends on the composition of both fluids and on the temperature. The
minimum miscibility pressure is referred to as the MMP. Laboratory experiments
(slim-tube experiments) may be performed for a particular oil and gas, at a particular
temperature to determine the MMP. See Figure 4.

When the pressure is lower than the MMP, gas injection improves recovery by
lowering the viscosity of oil, reducing the surface tension and also swelling the oil
(Brock et al, 1989). However, when miscibility occurs, the interfacial tension reduces
almost to zero, so the recovery may be very high (Figure 4). Some engineers specify
that MMP occurs when the recovery factor is 95% (in a slim tube).

Simulations of a slim-tube experiment using compositional simulation can show similar


results, using a very fine-scale 1D model (cell length ~ cms). However, the high
recoveries obtained in a slim-tube are not necessarily found in a real reservoir: the
effects of gravity and heterogeneity mean that the gas does not contact all the oil.

Figure 4
Example of a slim tube experiment, showing recovery of oil after the injection of 1.2
PV of CO2 as a function of pressure. (After Yellig and Metcalfe, 1980.)

Miscibility can be first contact miscible, when any amount of solvent (i.e. gas) forms a
single phase with the oil (Holm, 1986). One the other hand, miscibility can also be

©HERIOT-WATT UNIVERSITY December 2017 v2


9

multi-contact, or dynamic. There are two mechanisms which may take place:
vaporizing gas drive or condensing gas drive (Holm, 1986).

Of particular importance is CO2 injection. CO2 has been used for EOR in the USA, since
the 1970s. The CO2 comes from natural underground CO2 reservoirs, and is
transported to the reservoirs by a network of pipes (over 3000 miles of pipes). The
advantage of using CO2 compared with other gases is that the MMP for CO2 is lower.
Also, the density is higher, which means there is less gravity over-ride of the gas. Figure
5, from Holm and Josendal (1974) shows how different components of oil mix with CO2
gas, depending on whether the flood is near-miscible, multi-contact miscible or
miscible.

Slim-tube simulations can also demonstrate how CO2 reacts with different components
of the oil (Figure 6). (Note that the composition of the oil in this simulation was
different from that in Figure 5; we are not trying to compare “like with like”.) It is
interesting to see how the components have been split. Methane (C1), which has a
high mobility is concentrated in a “bank” at the head of the flood front. Also,
comparing the immiscible and miscible results, you can see a clear difference in the
remaining oil in the region swept by the CO2. Although Figure 6 shows a 2D distribution
of components, this is inferred from the results which contained mole fractions of the
different components in the oil and gas phases. The model was only a 1D model.
Unfortunately, this level of detail cannot be achieved when simulating CO2 injection at
the reservoir scale – a coarser-scale model is required, and numerical dispersion will
affect the results.

Figure 5
Experimental results of a CO2 flood at miscible and near-miscible conditions (based
on floods in a long slim tube at 135 oF. From Holm and Josendal (1974).

©HERIOT-WATT UNIVERSITY December 2017 v2


10

Figure 6
Example results of a fine-scale 1D simulation of CO2 flood (G. Wang, 2016,
unpublished).

Gas injection has two disadvantages, which reduce the sweep:


 Gas has a lower density than oil, so over-rides the oil in the reservoir;
 Gas has a low viscosity, and therefore a high mobility compared with oil, and this
means that the flood is unstable. Small heterogeneities in the rock will cause
irregularities in the front to grow forming “fingering”.

These problems are often overcome by using “WAG” which is water alternating gas.
In this procedure “slugs” of water and gas are injected into the reservoir. There are
two advantages of WAG:
 Since water is denser than oil it can help to recover oil at the bottom of the
reservoir, while gas being less dense can displace oil at the top of the reservoir;
 The microscopic sweep efficiency is increased, i.e. the residual oil at the pore-scale
is reduced.

Reservoir simulation is often used to investigate the “slug ratio” – i.e. the amount of
gas injected compared with water (each measured in terms of pore volume of the
reservoir), and also the slug size (length of time a slug is injected).

©HERIOT-WATT UNIVERSITY December 2017 v2


11

Figure 7
Schematic diagram of WAG, showing how it can improve the sweep.

2.3 CO2 Storage


In the USA, CO2 has been injected into the subsurface solely for EOR purposes.
However, recently it has been recognised that the burning of fossil fuels is increasing
the level of CO2 in the atmosphere and causing global warming. One solution to
mitigate this is to capture CO2 at power stations (or other industrial sites), transport it
to a suitable location, and inject into the subsurface (e.g. IPCC, 2005). CO2 may be
stored in abandoned oil and gas reservoirs, abandoned coals mines, or in deep saline
formations (aquifers). In addition, of course, CO2 can be injected into oil reservoirs for
CO2 EOR. Although, some of the CO2 will be produced with the oil, some CO2 will
remain stored in the reservoir. Deep saline aquifers provide the largest storage
capacity worldwide. The aim is to inject CO2 into formations at least 800 m deep,
where the pressure and temperature are high enough for CO2 to be supercritical – i.e.
it will have a liquid-like density (~ 700 kg/m3), but a gas-like viscosity (~ 0.05 cP). When
CO2 is injected into a deep saline formation, several process take place:
 The pressure increases in the formation;
 CO2 rises due to buoyancy;
 Some CO2 dissolves in water (brine);
 After injection has ceased, CO2 continues to rise, but water is imbibed into the CO2
plume, giving rise to pore-scale trapping of CO2;

©HERIOT-WATT UNIVERSITY December 2017 v2


12

 When CO2 dissolves in brine, a weak acid is formed which gives rise to chemical
changes in the rock.

Note that, since the CO2 rises, there must be an impermeable caprock above the
storage formation, just like in an oil or gas reservoir, to prevent CO2 from reaching the
surface.

All of these processes may be simulated using reservoir simulation software. Often
compositional simulation is used. However, sometimes black-oil simulators are used,
by pretending that water is oil and CO2 is hydrocarbon gas.

Reservoir simulation is important for assessing the fate of injected CO2. Companies
wishing to start a CO2 storage project must demonstrate to regulatory bodies and to
the general public that the CO2 will remain securely stored for thousands of years.

Figure 8
Example simulation of CO2 storage in a saline aquifer. CO2 was injected into a model
of a sandstone formation. The caprock is not shown, so that the distribution at the top
of the storage formation can be seen. The size of the model is approximately 40 km x
30 km x 300 m. CO2 was injected for 15 years, and then the well was shut in. The
simulation was continued for 1000 years to observe what would happen in the long
term.

Since simulations of CO2 injection are slow, especially in larger aquifers (which are
much bigger than oil reservoirs), some engineers have developed methods to speed
up simulation of CO2 injection. In particular, the movement of the CO2 plume is
dominated by buoyancy, so some engineers run simulations in vertical equilibrium
(VE). VE assumes that fluids separate out instantaneously so that the less dense fluid
(CO2) is on top of the more dense fluid (brine – i.e. saline water). This is the case for

©HERIOT-WATT UNIVERSITY December 2017 v2


13

the example shown in Figure 9 below, which is one of the examples in MRST. CO2 was
injected for 30 years, and then the simulation was continued for another 470 years.
(This example was run using a previous version of MRST. To run the same example in
the latest version requires downloading some extra routines.)

Figure 9
Example simulation of CO2 injection using the MRST co2lab module. This diagram
shows the 3D model at the top left, a view from above at the bottom left, two cross-
sections bottom right, and proportions of trapped and moveable CO2 top right (Lie
2016).

3. Chemical EOR
A variety of chemicals may be injected into a reservoir to improve recovery. As shown
in Figure 1, there are two mechanisms by which chemicals can increase oil recovery:
 By improving the mobility ratio which improves the macroscopic sweep
 By reducing interfacial tension (capillary pressure) to increase the microscopic
sweep.

Some EOR methods use a combination of chemicals so that both of the above
mechanisms take place.

©HERIOT-WATT UNIVERSITY December 2017 v2


14

3.1 Polymer Floods


Water is generally less viscous than oil, so it is more mobile, and waterfloods may be
unstable. This means that small heterogeneities in the formation can lead to early
breakthrough of water. First we look at fluid mobilities, which were mentioned in
Topic 5. Mobility is defined as relative permeability divided by viscosity:

krf
f  (5)
f

where f is fluid (o or w).


Then the mobility ratio for two fluids (e.g. oil and water) is defined as:

Mobility of displacing fluid w krw w krw o


M    . (6)
Mobility of displacedfluid o kro o kro w

If M < 1 a flood is stable, but if M > 1 it is unstable. Of course the relative permeability
varies with saturation. Usually M is evaluated using the end-point relative
permeabilities – i.e. the maximum values.

For example, if krw,max = 0.25, kro,max = 0.85, w = 1 and o = 3, then M = 0.88, which is
stable.
However, if krw,max = 0.25, kro,max = 0.85, w = 1 and o = 10, then M = 2.94, which is
unstable.

3.1.1 Buckley-Leverett Theory and Polymers


In practice, there are several effects that arise when polymer is injected into a
reservoir. For example, the polymer may be adsorbed onto the grains of rock, the
relative permeability to water may be reduced or the polymer viscosity may be
lowered due to shearing. These effects can be taken into account in commercial
numerical simulators. Here we look at a simple example to show the effect of
increasing the water viscosity in Buckley-Leverett displacement. You have already
performed similar calculations in Tutorials.

Figure 10
Relative permeability curves used for the 1D model.

©HERIOT-WATT UNIVERSITY December 2017 v2


15

Figure 11 shows some example relative permeabilities and


Figure 11 shows the resulting fractional flow curves for different viscosities. The solid
fw curve (red, in colour) is for a low viscosity oil (o = 2.5 cP) and water (w = 1.0 cP);
the fw curve with the short dashes (blue) is for a more viscous oil (o = 10.0 cP), and the
same water viscosity; thirdly, the fw curve with the long dashes (green) is also for the
more viscous oil, but the water viscosity has been increased (w = 8.0 cP) to represent
water with polymer. Compare the Welge tangents for these three curves (Topic 2,
Section 5), and then estimate what the flood profiles should look like.

From Figure 12 you can see that the case with the water flood with the heavy oil has a
lower shock front height and water breaks through earlier. On the other hand, the
case with the polymer flood has a high shock front height and breaks through later.

Note though, that in reality, fractional flow theory for the injection of polymer into a
reservoir is slightly more complex, because there will be a connate water “bank”
preceding the front of water with polymer.

Figure 10
Relative permeability curves used for the 1D model.

©HERIOT-WATT UNIVERSITY December 2017 v2


16

Figure 11
Fractional flows for different viscosities.

Figure 12
Saturation profiles for the examples above.

3.1.2 Numerical Simulation Example


Figure 13 shows examples of a water flood of a 2D model. The model has 15 layers,
and the top 5 layers have a higher permeability (1000 mD) than the rest of the model
(200 mD). The model initially contains oil (and connate water). There is a water
injector in the left column of cells and an oil producer on the right. Table 1 shows the
viscosities of the fluids used. It can be seen that the case with the polymer gives a
much more thorough sweep.

Fluid Viscosity (cP)


Light oil 1.74
Heavy oil 8.70
Water 0.80

©HERIOT-WATT UNIVERSITY December 2017 v2


17

Water + polymer 10.0

Table 1
Viscosities of fluids in the numerical example.

Figure 13
Waterfloods in a 2D model. Top: Oil with low viscosity. Middle: Oil with high
viscosity. Bottom: Oil with high viscosity, with polymer added to the water. See text
for more explanation.

4. Thermal EOR
As mentioned in the introduction, some reservoirs have heavy, very viscous oil. The
best way to recover this is to apply heat to reduce the viscosity. We give a very brief
overview of these methods here. Thermal EOR may be carried out in several ways (e.g.
Lake, 1989):
 Inject hot water
 Steam Soak
o Also known as “huff ‘n puff”. Steam is injected into a well, and then the well
is closed for a period time to allow the steam to warm the formation. Then
the well is put on production.
 Steam Drive

©HERIOT-WATT UNIVERSITY December 2017 v2


18

o This is like waterflood, except steam is injected. Sometimes a process called


SAGD is used – Steam Assisted Gravity Drainage which involves two horizontal
wells one above the other. The top well injects steam, and oil flows to the
lower producer. See Figure 14.
 In situ combustion
o Air or oxygen is injected and then ignited. Connate water is vaporized and so
this acts like a steam flood. Also some of the lighter components of the oil are
vaporised and these act like a miscible flood (Lake, 1989).

Commercial simulators like Eclipse and CMG have modules for simulating thermal EOR.
For example the CMG modelling group have a package known as STARS –Thermal &
Advanced Processes Reservoir Simulator.

Figure 14
Schematic diagram of SAGD. After Butler, 1994. The diagram is a vertical cross
section, looking along the horizontal wells, which appear as circles.

5. Simulation in Fractured Reservoirs

5.1 Introduction
At least half of the world’s remaining oil reserves are in fractured reservoirs,
particularly in carbonate reservoirs. Fractured reservoirs provide a challenge for
simulation due to the fact that fractures are very narrow, but have a very high
permeability. Nelson (2001) provided a classification of reservoirs:

Type I: Fractures provide essential porosity and permeability


Type II: fractures provide essential reservoir permeability

©HERIOT-WATT UNIVERSITY December 2017 v2


19

Type III: Fractures assist permeability in an already producible reservoir


Type IV: Fractures provide no additional porosity or permeability but create reservoir
anisotropy (barriers).

Figure 15
Nelson’s classification for different types of fractured reservoir. On the axes “k” =
permeability, “” = porosity, “m” = matrix and “f” = fractures. (After Nelson, 2001.)

Obviously the type of reservoir will influence the type of method which is used to
simulate flow in the reservoir. For Types 1 and 2, it is very important that the fractures
are carefully modelled in a simulation, whereas this is less important for Types 3 and
4. But, whatever type of reservoir is being studied, it is clear that it is not feasible to
make a grid and perform a simulation which has all of the fractures represented – the
model would be too complex and take too long to simulate. In addition, since the
fracture properties (length, permeability, connectivity) are uncertain, many versions of
a model should be assessed.

5.2 Dual Porosity Method


One early method to simplify models of fractured reservoirs was developed by Warren
and Root (1963), namely the Dual Porosity Method. This method is still used widely
today, and is available in commercial reservoir simulation packages. Warren and Root
(1963) assumed that the fractures were parallel to the main axes of the model (i.e.
along the x, y and z directions), and that there were regularly sized matrix blocks in
between (as though the model was made of sugar cubes). See Figure 16. They used
two grids to represent the system – one grid for the matrix and one for the fractures.
Most of the flow is through the fractures in the fracture grid, but fluids are also able to
flow between the matrix and the fractures via a transmissibility function (
Figure 17).

©HERIOT-WATT UNIVERSITY December 2017 v2


20

Figure 16
Left: fractured rock. Right: “sugar cube” model. (Warren and Root, 1963.)

Figure 17
Dual porosity grid. Fluid can flow between the fracture blocks, and also between the
matrix and fracture blocks via a transmissibility function.

The volumetric flowrate (Q) between the matrix and the fractures is given by:

Q  const  km   V   Pm  Pf  (7)

where km is the matrix permeability, V is the volume of a matrix block, Pm and Pf are
the pressures in the matrix and fracture grid cells, and  is a “shape factor” which
represents the area of the matrix in contact with the fractures divided by the size of
the matrix blocks. A simple formula for the shape factor was given by Kazemi (1976):

©HERIOT-WATT UNIVERSITY December 2017 v2


21

1 1 1
  4  2 2 (8)
 l 2
l y lz 
 x

where lx, ly and lz are the dimensions of the matrix blocks, assuming rectangular blocks.
Many researchers have improved on this equation. The value of  depends on the
shape of the matrix blocks and on the recovery mechanism – for example the
importance of gravity vs capillary forces. If a reservoir is being water flooded and the
matrix is water-wet, then water may be imbibed from the fractures into the matrix,
from all sides. On the other hand, if gravity is important, the flow between matrix and
fractures will occur in the vertical direction. Much research has been carried out on
determining appropriate formulae for the shape factor, but this is beyond the scope of
this course.

Note that the dual porosity approach is suitable for reservoir of Nelson types II and III,
where there is some permeability in the matrix, but the fracture permeability is more
important. In some cases, dual permeability simulations may be run where the matrix
blocks are also in communication with each other.

5.3 Discrete Fracture Models


Since dual porosity models were first developed, there has been much research into
the nature of fractures. More data has been gathered to give a greater understanding
of how fractures arise and how they are distributed, and so fractures can be modelled
in much greater detail (although there is still much uncertainty). Also, due to a huge
improvement in computer capabilities over the years, methods have been developed
to simulate fractures explicitly. The term “DFN” is used to denote discrete fracture
networks, where fractures are modelled as planes. An example is shown in Figure 18.)
In a reservoir where the matrix has negligible porosity and permeability (Nelson Type
I), flow could be simulated only in a fracture network. On the other hand when the
matrix does contribute to the flow, some engineers have developed sophisticated
methods (finite volume methods) to simulated flow through the fractures and the
matrix. “DFM” stands for discrete fracture and matrix (Ahmed Elfeel et al, 2012).

Fractures occur at all scales, ranging from tiny cracks to reservoir-scale fracture
corridors. It is common to take a staged approach to modelling fractured reservoirs.
A discrete fracture model may be generated at the small-scale and then upscaling
performed to create effective properties for a single-porosity (i.e. conventional) or a
dual-porosity model at the field scale. It is important to determine the REV
(Representative Elementary Volume – Topic 6, Section 2.1) of the small-scale fractures
so that appropriate scales are used for upscaling.

©HERIOT-WATT UNIVERSITY December 2017 v2


22

Figure 18
Discrete Fracture Model, showing two sets of fractures (green and blue). (Ahmed
Elfeel and Geiger, 2012).

Estimate of Fracture Permeability


Although fracture surfaces are actually rough, it is often assumed for simplicity the
fractures are smooth planes with a fixed aperture size, d. Figure 19 shows a set of
fractures in the x-z plane. The effective permeability for flow through a fracture in x
direction is:

d2
k . (9)
12

From Darcy’s Law, the volumetric flow rate through the fracture is:

kA P k yz P d 3z P
q   . (10)
 x  x 12 x

In Equation (10) above, we have used A = yz (area perpendicular to flow), and y =
d in a fracture. There the flow rate is proportional to the cube of the fracture width.
This is sometimes referred to as the cubic law.

Note that Equation (9) gives a simple way to estimate fracture permeability. For
example if the aperture is 10 m, then the permeability is:

105 105
k  8.33333 1012 m2
12
 8.33333 1012 9.869223 1013 D  8.44 D

©HERIOT-WATT UNIVERSITY December 2017 v2


23

Figure 19
Simple fracture model.

Therefore a very narrow fracture can have a very high permeability. If the matrix blocks
are of size 1 m cube and their permeability is 0.1 mD, this emphasizes the huge
contrasts of scale and permeability which arise in a fractured reservoir model.

Obviously, simulations through detailed small-scale models are time consuming, so an


analytical method is sometimes used. Oda (1985) developed an analytical method for
upscaling fracture models. This method uses Equation (9) to calculate fracture
permeability, and assumes that there is a set of fractures with a particular distribution
of fracture orientations, fracture sizes (surface area) and apertures and integrates to
obtain the effective permeability. It is available in some commercial packages, such as
Schlumberger’s Petrel.

Although Oda’s (1985) method is adequate for models with well-connected fractures
which are long compared with the grid block size, it is less accurate in other cases
compared to numerical flow simulation (Ahmed Elfeel and Geiger, 2012). In order to
simulate fractures explicitly, usually an unstructured grid (e.g. with triangular
elements, or tetrahedral in 3D) is used and a finite volume or finite element model is
required.

Some methods have been developed to simulate both fractures and matrix – DFM
models (e.g. Geiger et al, 2009). These are very large complex models, and they require
parallel processing to be able to complete the simulations within a reasonable time.
This means that the model is split into sections and different sections are simulated on
different processors (Figure 20). The advantage of simulating flow in both fractures
and matrix is that the simulator can handle a range of Nelson Types.

©HERIOT-WATT UNIVERSITY December 2017 v2


24

Figure 20
Complex Fracture-Matrix model. The different colours indicate how the model was
split up for simulation on different processors.

6. Closing Remarks
This Topic has covered a range of sub-topics to show more advanced uses of reservoir
simulation which you might come across in the future. The subject split into two parts:
simulation of enhanced oil recovery and simulation in fractured reservoirs.

On the subject of EOR, you should now be able to describe the following:
 Gas injection (miscible/immiscible)
 Compositional simulation
 CO2 injection for EOR and storage
 Polymer injection, and its effect on the flood front
 Thermal EOR for heavy oil reservoirs.

For fractured reservoirs you should be able to describe:


 The importance for simulating flow in fractured reservoirs
 The main issues which cause problems
 The dual porosity model
 Explicit modelling of fractures.

©HERIOT-WATT UNIVERSITY December 2017 v2


25

7. References
Ahmed Elfeel, M. and Geiger, S., 2012. “Static and Dynamic Assessment of DFN
Permeability Upscaling”, SPE 154369, presented at the EAGE Annual Conference and
Exhibition, incorporating SPE Europec, Copenhagen, Denmark, 4-7 June, 2012.

Brock, W.R. and Bryan, L.A., 1989. “Summary Results of CO2 EOR Field Tests, 1972 –
1987”, SPE 18977. Presented at the SPE Joint Rocky Mountain Regional/Low
Permeability Reservoirs Symposium and Exhibition, Denver, Colorado, 6-8 March,
1989.

Butler, R.M., 1994. “Steam-Assisted Gravity Drainage: Concept, Development


Performance and Future”, Petroleum Society of Canada. doi:10.2118/94-02-05.

Geiger, S., Huangfu, Q, Reid, F., Matthai, S., Coumou, D., Belayneh, M., Fricke, C. and
Schmid, K., 2009. “Massively Parallel Sector Scale Discrete Fracture and Matrix
Simulations”, SPE 118924, presented at the Reservoir Simulation Symposium, The
Woodlands, Texas, 2-4 February, 2009.

Holm, L.W., 1986. “Miscibility and miscible displacement”, Journal of Petroleum


Technology, 38 (8), 817-818.

Holm, L.W. and Josendal, V.A., 1974. “Mechanisms of oil displacement by carbon
dioxide”, Journal of Petroleum Technology, 26 (12), 1427-1438.

IPCC, 2005. “Carbon Dioxide Capture and Storage”, Special Report, Chapter 5:
“Underground Geological Storage”,
http://www.ipcc.ch/publications_and_data/publications_and_data_reports.htm#1.

Kazemi, H., Merrill Jr, L.S., Porterfield, K.L. and Zeman, P.R., 1976. “Numerical
Simulalion of Water-Oil Flow in Naturally Fractured Reservoirs”, SPE 5719, SPE
Journal, 16 (6), 317 – 326.

Lake, L.W., 1989. “Enhanced Oil Recovery”, Prentice Hall Inc, New Jersey.

Lie, K-A, 2016. “An Introduction to Reservoir Simulation Using MATLAB: User Guide
for the Matlab Reservoir Simulation Toolbox (MRST)”, www.sintef.no/mrst.

Warren, J.E. and Root, P.J., 1963. “The Behavior of Naturally Fractured Reservoirs”,
SPE 426, SPE Journal, 3(03), 245-255.

©HERIOT-WATT UNIVERSITY December 2017 v2


26

Yellig, W.F. and Metcalfe, R.S., 1980. “Determination and Prediction of CO2 Minimum
Miscibility Pressures”, Journal of Petroleum Technology, 32 (1), 160-168.

©HERIOT-WATT UNIVERSITY December 2017 v2

Das könnte Ihnen auch gefallen