The Imperial CollegeLecturesinPetroleumEngineeringVolume 3
The Imperial CollegeLecturesinPetroleumEngineeringVolume 3
The Imperial CollegeLecturesinPetroleumEngineeringVolume 3
PETROLEUM ENGINEERING
Topics in Reservoir Management
Volume
Volume
Deryck Bond
Kuwait Oil Company, Kuwait
Samuel Krevor
Imperial College London, UK
Ann Muggeridge
Imperial College London, UK
David Waldren
Petroleum Consulting and Training (PCT) Ltd, UK
Robert Zimmerman
Imperial College London, UK
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
Preface
This book is the third volume of a set of lecture notes based on the
Master of Science course in Petroleum Engineering that is taught
within the Department of Earth Science and Engineering at Imperial
College London. The Petroleum Engineering MSc is a one-year course
that comprises three components: (a) a set of lectures on the different
topics that constitute the field of petroleum engineering, along with
associated homework assignments and examinations; (b) a group field
project in which the class is broken up into groups of about six
students, who then use data from an actual reservoir to develop the
field from the initial appraisal based on seismic and geological data,
all the way through to eventual abandonment; and (c) a 14-week
individual project, in which each student investigates a specific
problem and writes a small “thesis” in the format of an SPE paper.
The Petroleum Engineering MSc course has been taught at
Imperial College since 1976, and has trained over a thousand
petroleum engineers. The course is essentially a “conversion course”
that aims to take students who have an undergraduate degree in
some area of engineering or physical science, but not necessarily
any specific experience in petroleum engineering, and train them
to the point at which they can enter the oil and gas industry as
petroleum engineers. Although the incoming cohort has included
students with undergraduate degrees in fields as varied as physics,
mathematics, geology, and electrical engineering, the “typical” stu-
dent on the course has an undergraduate degree in chemical or
v
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-fm page vi
Robert W. Zimmerman
Imperial College London
July 2017
August 23, 2017 14:34 Topics in Reservoir Management - 9in x 6in b2764-fm page vii
Deryck Bond holds BSc and PhD degrees in Physics from Imperial
College London. He has worked in research, reservoir engineering and
petroleum engineering functions for a number of major and mid-sized
oil companies and as an independent consultant. He has contributed
to the Petroleum Engineering MSc program at Imperial College. He
is currently a Senior Consultant working on the Greater Burhan field
for Kuwait Oil Company.
vii
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-fm page viii
Contents
Preface v
About the Authors vii
ix
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-fm page x
Contents xi
Contents xiii
Contents xv
4.5.6
Outline of Approach to Matching the Model . . 223
4.5.7
How Well Should We Aim to Match Data? . . . 224
4.5.8
Assessing the “Goodness” of a Match . . . . . . 226
4.5.9
Well Controls During the History Match . . . . 227
4.5.10
Initial Simulation Runs . . . . . . . . . . . . . . . . 229
4.5.11
Review of Scope for Changing Model Input
Parameters . . . . . . . . . . . . . . . . . . . . . . . 230
4.5.12 Sensitivity Study . . . . . . . . . . . . . . . . . . . . 231
4.5.13 The History Matching Process . . . . . . . . . . . 232
4.5.14 Matching Pressure . . . . . . . . . . . . . . . . . . . 233
4.5.14.1 Pressure match example . . . . . . . . . 234
4.5.15 Matching Fluid Movement . . . . . . . . . . . . . 236
4.5.16 Matching Water Movement — Example 1 . . . 238
4.5.17 Matching Water Movement — Example 2 . . . 240
4.5.18 Matching Water Movement — Example 3 . . . 241
4.5.19 Matching Well Pressures/Detailed Well
Performance . . . . . . . . . . . . . . . . . . . . . .. 242
4.5.20 The Transition to Prediction . . . . . . . . . . .. 243
4.6 Automatic and Computer-Assisted History
Matching/Multiple Matches . . . . . . . . . . . . . . . . .. 243
4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . .. 243
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 244
Index 245
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 1
Chapter 1
Robert W. Zimmerman
Department of Earth Science and Engineering,
Imperial College London, Kensington, London SW7 2AZ, UK
1.1. Introduction
The most important fact about reservoir rocks is that, by definition,
they are not completely solid, but rather are porous to one degree
or another. The degree to which they are porous is quantified by a
parameter known as the porosity. The fact that the rocks are porous
allows them to hold fluid. If these pores are interconnected, which
they are in most rocks, then the fluid is able to flow through the
rock, and the rock is said to be permeable. The ability of a rock to
allow fluid to flow through it is quantified by a parameter called the
permeability. As the porosity controls the amount of oil or gas that
the rock can hold, and the permeability controls the rate at which
this oil or gas can flow to a well, these two parameters, porosity
and permeability, are the most important attributes of a rock, for
reservoir engineering purposes.
It would be very advantageous to petroleum engineers if the pore
space of a reservoir were completely filled with hydrocarbon fluid.
Unfortunately, this is never the case, and the pores always contain
a mixture of hydrocarbons and water. The relative amounts of oil,
gas or water are quantified in terms of parameters known as the fluid
saturations. These saturations are in turn controlled by the surface
interactions between the rock and the various fluids, which can be
1
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 2
Sand grains
Pore
Figure 1.2.1. Schematic diagram of porous sandstone, showing grains and pore
space. Typical grain sizes are tens to hundreds of microns.
φ = Vp /Vb . (1.2.2)
and effective porosity, which measures only the pore space that is
interconnected and which can potentially form a flow path for the
hydrocarbons. The total porosity is therefore composed of effec-
tive/interconnected porosity, and ineffective/unconnected porosity.
In most rocks, there is little ineffective porosity. One important
exception are carbonate rocks called diatomites, in which most
of the porosity is unconnected. The Belridge oilfield in central
California has produced 1.5 billion barrels of oil from a diatomite
reservoir that has porosities ranging from 45% to 75%, most of
which is unconnected and not “effective”! But this is a special case
that requires special production methods, and is not typical of the
reservoirs that will be the focus of most of this course.
Sand
10 m
Shale
x1
x2
(a) (b)
0 R
REV
Figure 1.2.4. Porosity as a function of the size of the sampling region, showing
the existence of an REV.
1.2.3. Saturation
Reservoir rocks are never filled completely with oil, for reasons that
will be discussed later in this chapter. Consider a rock that contains
some oil and some water, as in Fig. 1.2.5. If the volume of water
contained in a region of rock is Vw , the volume of oil is Vo , and the
total volume of the pore space is Vp , then the saturation of each
of these phases can be defined as the fraction of pore space that is
occupied by that phase, i.e.
Sand
Oil grain
Water Pore
Figure 1.2.5. Schematic diagram of a porous rock containing oil and water.
Sw + So = 1. (1.2.4)
Sw + So + Sg = 1. (1.2.5)
kA(Pi − Po )
Q= , (1.3.1)
µL
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 8
Pi
Po
L
x
Q
A
Figure 1.3.1. Experimental setup for measuring the permeability of a porous rock
or sand.
where
Q k(Pi − Po )
q= = , (1.3.2)
A µL
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 9
where the flux q has dimensions of m/s. Please note that the flux
is not the same as the velocity of the fluid particles,1 and so it is
perhaps easier to think of these units as m3 /m2 ·s.
For the general case in which the flux may vary from point-
to-point, we need a differential form of Darcy’s law. The differential
version of Eq. (1.3.2) for horizontal flow is,
−k dP
qx = . (1.3.3)
µ dx
The minus sign is included to account for the fact that the fluid flows
in the direction from higher to lower pressure.
For vertical flow, we must include a gravitational term in Darcy’s
law. To see why this is necessary, recall from fluid mechanics that if
the fluid is stagnant, then the pressure distribution will be
P = Po + ρgz, (1.3.4)
where z is the depth below some datum level, and Po is the pressure
at the datum level. So, there will be pressure gradient in a stagnant
fluid, but there will be no flow. The “equilibrium pressure gradient”
is, from Eq. (1.3.4),
dP
= ρg. (1.3.5)
dz equilibrium
1
Note: q is the flux based on the total nominal area of the core, but the fluid
actually flows only through the pores, and not the grains! So, the total flux is
given by Q = qA, but it can also be expressed as Q = vApore , where v is the
actual mean velocity of the particles of fluid, and Apore is area occupied by pores.
Hence, qA = vApore , so v = q(A/Apore ) = q/φ. For example, if q is 1 cm/h in a
reservoir of 10% porosity, the actual mean velocity of the oil molecules as they
travel through the pore space is 10 cm/h.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 10
Φ = P − ρgz, (1.3.8)
2
Caution: if the rock is anisotropic, then Eq. (1.3.9) does not hold in an arbitrary
direction, even if we use an appropriate value of k (de Marsily, 1986). The correct
version of Darcy’s law for an anisotropic rock must be written in terms of the
permeability tensor, which is a symmetric 3 × 3 matrix that has six independent
components. However, this tensorial form of Darcy’s law is not typically used in
most reservoir engineering calculations.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 11
d L
π d4 ∆P
Q= . (1.3.13)
128µ L
If there are N such pores, the total flowrate will be
N π d4 ∆P
Q= . (1.3.14)
128µ L
The total area of these pores, in the plane of the page, is Ap =
N π d2 /4, and the porosity is φ = Ap /A = Ap /L2 , where A is the
macroscopic area normal to the flow. Hence, the total flowrate from
Eq. (1.3.14) can be written as
φd2 A ∆P
Q= . (1.3.15)
32µ L
φd2
k= . (1.3.16)
96
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 14
∆Φ
Q1 k1 H1
Q2 k2 H2
.....
QN kN HN
Fig. 1.3.3. Within each layer, fluid will flow horizontally, according
to Darcy’s law:
−ki (Hi w) ∆Φ
Qi = , (1.3.18)
µ ∆x
where w is the thickness into the page.
The total flowrate is found by summing up the flowrates through
each layer:
N
N
N
−ki (Hi w) ∆Φ −w ∆Φ
Q= Qi = = ki Hi , (1.3.19)
µ ∆x µ ∆x
i=1 i=1 i=1
Hence, the effective permeability for flow along the layering is the
weighted arithmetic mean of the individual permeabilities, weighted
by the thickness of the layers.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 16
Q
Area, A
k1 H1
k2 H2
∆Φ
.....
kN yN
Figure 1.3.4. Fluid flow perpendicular to the layering through a layered rock.
−ki A ∆Φi
Qi = , (1.3.22)
µ Hi
where A is the area normal to the flow direction, i.e. in the horizontal
plane.
In the steady state, the flowrate through each layer must be the
same, but the potential drop will be different. So, we put Qi = Q in
each layer, and rewrite Eq. (1.3.22) in the form
−µQHi
∆Φi = . (1.3.23)
Aki
N
N
N
−µQHi −µQ Hi
∆Φ = ∆Φi = = . (1.3.24)
Aki A ki
i=1 i=1 i=1
The expression on the right side of Eq. (1.3.26) is called the weighted
harmonic mean of the permeabilities.
Equations (1.3.21) and (1.3.26) are similar to the equations for
the overall conductivity of electrical resistors in parallel or series.
However, this analogy can easily be remembered incorrectly, because
the thickness also appears in these equations, and it appears in a
different way in the two cases. Rather than trying to remember the
analogy between electrical circuits and flow through layered rocks,
it is safer to derive the laws for the effective permeability from first
principles (or to refer to these notes).
Roughly speaking, the effective permeability for flow parallel to
the layering is controlled by the permeability of the most permeable
layer, whereas for flow transverse to the layering, the least permeable
layer plays the controlling role.
Interface
FLL FLL
FLL FLL
Gas Liquid
U = T S − P V + γA, (1.4.1)
Wire frame
L
b F
Loop
dU = dW = FdL. (1.4.3)
dU = γdA. (1.4.4)
dU = γbdL. (1.4.5)
PL
PG 2R
added to the system, and no work is done on the system, so the total
entropy and energy remain the same). Hence,
which is equivalent to
PG − PL = 2γ/R. (1.4.11)
Gas
α Liquid
Solid
γLG
Gas Liquid
γGS α γLS
Solid
Figure 1.4.5. Force balance on the contact between three phases (liquid sitting
on a solid surface, surrounded by a gaseous phase).
In this case, we say that the liquid “wets” the solid surface, and
the surface is called “water-wet” (although a better name would
be “water-wettable”, since a “water-wet” surface can be completely
dry!).
Case 2: γLG < γGS − γLS < 0:
In this case, the interfacial energy of a gas–solid interface is less than
that of the liquid–solid interface, so the solid will “prefer” to be in
contact with the gas. The right-hand side of Eq. (1.4.13) will lie
between −1 and 0, and so α will lie in the range
In this case, we say that the liquid does not wet the surface; it
will sit on the surface in a bubble-shape, with very little interfacial
contact between the liquid and the solid surface; see Fig. 1.4.6.
Case 3: |γGS − γLS | > γLG :
In this case, the right-hand side of Eq. (1.4.13) does not lie between
−1 and +1, and so there is no value of α that will satisfy the equation
of mechanical equilibrium! To see what will happen in this case,
consider the limiting case in which γGS − γLS = γLG , in which case
Gas α
α
Solid
Figure 1.4.6. A wetting liquid (left) and a non-wetting liquid (right) on a solid
surface.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 25
cosα = 1, and α = 0◦ . In this case, the liquid will spread out over
the surface, creating as much liquid–solid interfacial area as possible
(example: oil on water!). If γGS −γLS > γLG , then the same situation
will occur, and the liquid will continue to flow until it forms a thin
layer on the solid surface. By a similar argument, if γGS − γLS <
−γLG , then the gas will spread out to cover as much of the solid
surface as possible.
Most reservoir rocks are preferentially “water-wet” as opposed
to “oil-wet”. If a water-wet rock is partially saturated with oil and
water, the pore walls will “prefer” to be in contact with water rather
than with oil, and so the oil will tend to exist in the form of blobs,
as in Fig. 1.4.6.
T
Air Air θ P
A
D z
Oil Oil
Water Water
C P
Figure 1.4.7. (a) Oil, air and water in a bucket; oil lighter than water. (b) If a
capillary tube is inserted into the water, the water rises in the tube to a higher
level. (c) Force balance on the column of water in the tube.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 26
water will rise up in the tube by some height, h, above the original
oil–water contact.
We now calculate the capillary rise, h, by performing a vertical
force balance on the column of water in the tube, as in Fig. 1.4.7(c).
The bottom of the column is pushed upwards by the pressure in the
water at level C, acting over an area of πR2 . With the sign convention
that the z-axis decreases with depth, this force is −Pw (zC )πR2 .
At the top of the column is a downwards-acting force due to the
pressure in the oil at level zD ; this force is +Po (zD )πR2 . The surface
tension exerts an upwards force along the entire wetted perimeter of
the tube; its magnitude is T = 2πγow R. It acts at an angle θ to the
vertical, so its vertical component is −2πγow R cos θ. Finally, gravity
acts downwards on the column of water with a force
W = mg = ρw Vg = ρw πR2 (zC − zD )g. (1.4.16)
Summing all the vertical forces to zero gives
−Pw (zC )πR2 + Po (zD )πR2 − 2πγR cos θ + ρw πR2 (zC − zD )g = 0.
(1.4.17)
The pressure in the oil at location D is equal to atmospheric
pressure plus the pressure due to a column of oil of height (zD − zA ),
i.e. Po (zD ) = Patm +ρo g(zD −zA ). Similarly, we can see that Pw (zC ) =
Patm + ρo g(zB − zA ) +ρw g(zC − zB ). Inserting these expressions into
Eq. (1.4.17) gives
−[Patm + ρo g(zB − zA ) + ρw g(zC − zB )]πR2
+ [Patm + ρo g(zD − zA )]πR2
− 2πγow R cos θ + ρw πR2 (zC − zD )g = 0, (1.4.18)
which can be solved to find the capillary rise, h:
2γow cos θ
zB − zD = h = . (1.4.19)
(ρw − ρo )gR
Hence, the height to which water would rise in a tube of radius R is
proportional to the surface tension between the water and oil, and is
inversely proportional to the radius of the tube.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 27
Po (zD ) − Pw (zD )
= Patm + ρo g(zD − zA ) − Patm − ρo g(zB − zA ) + ρw g(zB − zD )
= (ρw − ρo )g(zB − zD ) = (ρw − ρo )gh = 2γow cos θ/R. (1.4.20)
In other words, Pcap = 2γow cos θ/R. This is identical to the Young–
Laplace equation that we derived earlier for an oil bubble in water,
modified to account for the contact angle. Note also that the capillary
pressure at any height h is equal to (ρw − ρo )gh. The theory we have
just described is referred to as “capillary-gravity equilibrium”.
B
Oil
Water
Figure 1.4.8. Similar to Fig. 1.16(b), but with a set of capillary tubes of different
radii. According to Eq. (1.4.19), the water will rise higher in the smaller tubes.
Oil/water contact
Free water level
0 Swi 1
Sw
Figure 1.4.9. Capillary pressure (left scale) and height above the FWL (right
scale), as functions of water saturation.
and so the y-axis in Fig. 1.4.9 essentially represents both the capillary
pressure and the height above the FWL, which differ only by the
multiplicative factor (ρw − ρo )g. Hence, this graph represents the
water–oil transition zone, but also represents the capillary pressure
function as a function of saturation.
Note that the relation Pcap = (ρw − ρo )gh holds regardless of
the specific rock geometry, assuming only that the rock is water-wet.
However, the precise form of the Pcap (Sw ) curve shown in Fig. 1.4.9
depends on the pore geometry, and specifically on the pore-size
distribution.
For the simple bundle-of-parallel-tubes model, one can derive
an exact relationship between the Pcap (Sw ) curve and the pore-size
distribution. For a real rock, in which the pores are interconnected,
the relationship is not so simple, but it is always true that a
narrow pore-size distribution corresponds to a Pc curve with a nearly
horizontal shape, whereas a broader pore-size distribution yields a
curve that increases more gradually, as shown in Fig. 1.4.10. The
extreme case of a bundle of tubes in which all pores had the same
Broad pore-size
distribution
Pcap
Narrow pore-size
distribution
0 Swi Sw 1
Figure 1.4.10. Capillary pressure curves for two rocks, with a narrow and a broad
pore-size distribution, respectively.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 30
1 k 1
Pcap = √ . (1.4.22)
γ cos θ φ 6
in the form
1 k
Pcap = J(Sw ). (1.4.23)
γ cos θ φ
1 k lab
P = J(Sw ). (1.4.24)
γlab cos θlab φ cap
1 k res
P = J(Sw ). (1.4.25)
γres cos θres φ cap
kro
krw
0
0 Siw 1-Sor 1
Sw
These two values, Siw and Sor , are also known as the relative
permeability end-points, and the relative permeability values at these
saturations are known as the end-point relative permeabilities.
The precise shapes of the relative permeability curves depend on
the details of the pore structure of the rock. Power-law functions are
often found to be useful in fitting these curves. Note that relative
permeability curves are never linear functions of the saturation,
although this simple linear form is sometimes assumed, particularly
in fractured reservoirs, as data on the relative permeabilities of the
fractures is rarely available. The range of validity of the assumption
of linear relative permeability functions in fractured reservoirs has
been investigated by de la Porte et al. (2005).
oil
water
(a)
oil
water
(b)
oil
water
oil
(c)
Figure 1.5.2. A pore doublet used to illustrate how small isolated blobs of oil can
get trapped behind when water displaces oil in a reservoir, giving rise to a finite
value of the residual oil saturation.
∆V
I= . (1.6.1)
R
Electrical charge has units of coulombs, so current, which is the flow
of charge, has units of coulombs/second. Resistance therefore has
units of volt-seconds/coulomb, which are also known as ohms. In the
form of Ohm’s law given by Eq. (1.6.1), R will depend on the material
properties, but also on the shape and size of the conductor.
Now consider a cylindrically shaped conductor, of length L and
cross-sectional area A, as in Fig. 1.6.1. All other factors being equal,
the current will be proportional to A, and inversely proportional to L.
L
A
∆V
Figure 1.6.1. Cylindrical core with cross-sectional area A, length L, and axial
voltage drop ∆V . If the material has conductivity σ, then the current will be
given by I = σA∆V /L.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 37
A1
A2
A3
Figure 1.6.2. Cylindrical tube model of a porous rock, used in developing a simple
model for the electrical formation factor.
F = 3φ−1 . (1.6.10)
F = bφ−m . (1.6.11)
1000
100
F F = 0.496φ−2.05
10
1
1 10 100
Porosity, φ (%)
electrical double layer then builds up along the surface of these plates.
This surface layer allows another path for current flow, separate
from the current flow through the brine-saturated pores that were
discussed above. To a good approximation, this surface current can
be thought of as being in parallel with the pore-current, and so
it adds an extra component to the conductivity of the rock. This
extra conductivity depends on the electrochemical properties of the
clays, but not on the intrinsic conductivity of the brine. The resulting
generalisation of Eq. (1.6.11) for shaly sands is
1
σ(brine-saturated rock) = (σw + BQv ), (1.6.13)
F
where Qv is the charge on the double layer, per unit volume, and B is
a constant. Equation (1.6.13), which is called the Waxman–Smits
equation, will be discussed further in the module on core analysis.
3
Note: when oil is taken from its pressurised state in the reservoir up to the
surface where it is at atmospheric pressure, any gas that had been dissolved in the
oil will be released, and the oil will actually shrink ! This phenomenon cannot be
described in terms of the compressibility of liquid oil; this shrinkage must be taken
into account when doing material balance calculations on a reservoir. However,
Eqs. (1.7.1) and (1.7.2) are applicable to the incremental pressure changes that
occur to the oil when it is flowing inside the reservoir.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 43
Pc
Pc
Pp
Pc
is defined as
−1 ∂V
C= . (1.7.3)
V ∂P
For a porous rock (Fig. 1.7.1), we need to consider two volumes,
the pore volume and the bulk volume, and two pressures, the pore
pressure and the confining pressure. So, we can define four different
compressibilities (Zimmerman, 1991):
−1 ∂Vb 1 ∂Vb
Cbc = , Cbp = , (1.7.4)
Vb ∂Pc Pp Vb ∂Pp Pc
−1 ∂Vp 1 ∂Vp
Cpc = , Cpp = . (1.7.5)
Vp ∂Pc Pp Vp ∂Pp Pc
References
Archie, G. E. (1942). The electrical resistivity log as an aid in determining
some reservoir characteristics, Petrol. Trans. AIME, 146, 54–62.
Bear, J. (1972). Dynamics of Fluids in Porous Media, American Elsevier,
New York.
Beran, M. (1968). Statistical Continuum Theories, Interscience, London.
de la Porte, J. J., Kossack, C. A. and Zimmerman, R. W. (2005). The
effect of fracture relative permeabilities and capillary pressures on the
numerical simulation of naturally fractured reservoirs. SPE Annual
Technical Conference held in Dallas, (SPE 95241).
de Marsily, G. (1986). Quantitative Hydrogeology, Academic Press,
San Diego.
Dullien, F. A. L. (1992). Porous Media: Fluid Transport and Pore Structure,
2nd ed., Academic Press, San Diego.
Leverett, M. C. (1941). Capillary behaviour in porous solids, Petrol. Trans.
AIME, 142, 159–172.
Mavko, G., Mukerji, T. and Dvorkin, J. (2009). The Rock Physics Handbook,
2nd ed., Cambridge University Press, Cambridge.
Ruffet, C., Gueguen, Y. and Darot, M. (1991). Complex conductivity
measurements and fractal nature of porosity, Geophysics, 56(6), 758–768.
Scheidegger, A. E. (1974). The Physics of Flow through Porous Media,
University of Toronto Press, Toronto.
Zimmerman, R. W. (1991). Compressibility of Sandstones, Elsevier,
Amsterdam.
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 45
Questions
1. Consider a reservoir that is shaped like a circular disk, 10 m thick,
and with a 5 km radius in the horizontal plane. The mean porosity
of the reservoir is 15%, the water saturation is 0.3, and the oil
saturation is 0.7.
(a) Ignoring the expansion of the oil that would occur when it is
produced from the reservoir, how many barrels of oil are in
this reservoir? One barrel = 0.1589 m3 .
(b) If the density of the oil is 900 kg/m3 , how much oil (in kg) is
contained in the reservoir?
2. In a laboratory experiment, a pressure drop of 100 kPa is imposed
along a core that has length of 10 cm, and a radius of 2 cm. The
permeability of the core is 200 mD, its porosity is 15%, and the
viscosity of water is 0.001 Pa·s.
(a) What will be the volumetric flowrate Q of the water, in m3 /s?
(b) What is the numerical value of q = Q/A, in m/s?
3. Imagine that the rock in problem 2 can be represented by the
parallel-tube model.
(a) Estimate the mean pore diameter, d, using Eq. (1.3.16).
(b) What is the mean velocity, v, of the water particles in the
rock?
(c) The importance of the inertia term, relative to the pressure
term in, say, Eq. (1.3.11), can be quantified by the Reynolds
number, defined as Re = ρvd/µ. What is the Reynolds number
in this experiment? Note that Darcy’s law is only accurate
when Re < 1 (Bear, 1972).
4. Consider a layered reservoir consisting of alternating layers, 1 m
thick, of rock 1, rock 2 and rock 3, where k1 = 1000 mD, k2 =
100 mD and k3 = 10 mD.
(a) What is the effective permeability of this rock, if fluid is
flowing parallel to the layering?
(b) What is the effective permeability of this rock, if fluid is
flowing perpendicular to the layering?
August 11, 2017 10:19 Topics in Reservoir Management - 9in x 6in b2764-ch01 page 46
(c) Imagine that the reservoir consists of these three rock types,
in equal volumetric proportions, but occurring in a “random”
spatial distribution. Estimate the effective permeability in
this case.
5. Consider a small blob of oil surrounded by water. The surface
tension between the oil and water is 0.02 N·m. If the radius of the
blob is 0.05 mm, what is the value of the capillary pressure? Is the
pressure higher in the oil or the water?
6. Consider again the parallel tube model of a rock. Assume that the
diameter of each pore is 20 mm, γ = 0.02N·m, ρo = 900 kg/m3 ,
ρw = 1000 kg/m3 , the contact angle θ is 45◦ , and g = 9.8 m/s2 . If
this rock is placed in a tank containing water overlain by oil, as
in Fig. 1.4.7, to what height will the water rise in the pores?
7. Consider a homogeneous reservoir with φ = 0.20, k = 200 mD,
water–oil surface tension of γ = 0.03 N/m, and oil–water contact
angle of 35◦ . The oil density is 850 kg/m3 , and the water density
is 1050 kg/m3 . In the lab, we determine that the irreducible water
saturation occurs when the J-function is equal to 4.23. What will
be the height of the oil–water transition zone in the reservoir?
Hint: Use Eq. (1.4.23) to convert J(Swi ) to Pcap , and use
Eq. (1.4.21) on to convert Pcap to height.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 47
Chapter 2
47
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 48
Primary production
• Natural lift
• Artificial lift
Secondary production
• Waterflooding
• Pressure maintenance
EOR
• Thermal stimulation
• Chemical EOR
• Gas injection processes
Improved oil recovery
• Includes EOR
• New wells
• Reservoir management to increase sweep
Unconventional
• Shale oil — light tight oil
• Extra heavy crude
• Kerogen (oil shale)
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 49
Canada: 21%
Indonesia:
11% Brazil, Norway,
Venezuela: 22% Gas injection:
Turkey, UK, Thermal: 55%
Trinidad, 45%
Germany and
USA: 42% Egypt 4%
× 10 5
Enhanced 8
production
(barrels per day)
6 Gas injection
4
Thermal
0
1992 1996 2000 2004 2008 2012
Figure 2.1.2. Oil produced using EOR techniques in the USA by technology since
1992.
Source: Koottungal (2014).
E. Europe/Eurasia
Africa 2012
2020
OECD Europe 2035
Middle East
Latin America
Asia Oceania
North America
0.2 0.4 0.6 0.8 1.0 1.2
(million barrels per day)
Figure 2.1.3. Modelled projections of oil production from all EOR operations
globally.
Source: (IEA, 2013b).
Base scenario
Figure 2.1.4. Modelled projections of oil produced from CO2 -EOR in the USA.
Source: (USEIA, 2014).
Other
Shale oil
conventional
Already MENA
produced conventional CO2-EOR
–
Technically recoverable resources (billion barrels)
Figure 2.1.5. Technically recoverable oil resources and associated production costs
in the absence of a CO2 price (non-CO2 -EOR) and in the presence of tax of
$150/tCO2 (CO2 -EOR).
Source: (IEA, 2013a).
RF = EP S × ES × ED × EC . (2.2.2)
vµ
Nc = , (2.2.3)
σ
where v is the the linear flow velocity, µ is the displacing fluid vis-
cosity and σ is the interfacial tension between the fluid phases. This
represents the balance between interfacial and viscous forces at the
pore scale and interfacial forces generally dominate in consolidated
rocks when Nc < 10−5 .
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 56
AI-Mansoori et al. (2009) Crowell et al. (1966) Geffen et al. (1952) Jerauld (1997)
Kleppe et al. (1997) Kralik et al. (2000) Land (Berea) (1971) Land (Alundum) (1971)
Ma and Youngren (1994) McKay (1974) Pickell et al. (1966) Suzanne et al. (2003)
70
60
50
S(nw)r (%)
40
30
20
10
0
0 10 20 30 40 50 60 70 80 90 100
S(nw)i (%)
single pore volume of water has swept through the system (Salathiel,
1973; Anderson, 1987).
The presence of hydrocarbon alters the wettability of the local
mineral surface to be oil wetting. Because the initially non-wetting
phase will have originally inhabited the larger pores, this often leads
to a scenario where the rocks are oil wet in the larger pores and water-
wet in the smaller pores. This is referred to as a mixed wet system and
it is thought to be the prevailing scenario in oil reservoirs. In mixed
wet systems nearly all of the oil can be displaced but this requires
many more pore volumes of a displacing, generally wetting, fluid to
be injected than production from water-wet systems (Salathiel, 1973;
Anderson, 1987).
The residual saturation is stable for a given rock so long as
capillarity dominates the pore scale force balance of the fluids in the
pores, i.e. Nc < 10−5 . A number of EOR techniques are understood
to work primarily through the disruption of the role of capillarity
in determining the local fluid distribution (Fig. 2.2.2). So-called
chemical techniques use aqueous fluids for the displacement of oil
but with dissolved solutes that dramatically lower the interfacial
mbr
oo
ste
Sor /Sor, wf
re
r
Ab loba
owsk
Du
&
ra
Pre
S
y
s
i&
0.5
Tab
d
Bro
er
Du P
wn
el l
rey
Nonwetting Residual
Wetting Residual
0
10–8 10–6 10–4 10–2 100
Capillary Number, Nc
tension between the water and the oil. Solvent flooding uses a
displacement fluid that is miscible with the hydrocarbon phase, such
as CO2 or other hydrocarbons, thereby picking up the residual oil
and recovering it with the injected solvent.
An approximate analysis can provide an estimate of how much
the interfacial tension needs to be reduced to result in an effect.
Typical interstitial field velocities are v ≈ 1 m d−1 ≈ 10−5 m s−1 .
The interfacial tension between oil and water is generally σ =
6 × 10−2 N m−1 and the viscosity of light crude oil is similar to
water, µ = 10−3 Pa·s. This results in Nc ≈ 10−7 which suggests
that increases of a factor of 100 or more are required to initiate
desaturation of the residual non-wetting phase. Thus EOR techniques
focused on reducing interfacial tension require chemicals that can
bring interfacial tension to less than 10−3 N m−1 . It is also worth
noting that a displacement at these capillary numbers also has
a large impact on the other multi-phase flow properties. Relative
permeability curves begin to increase and capillarity significantly
weakens (Fig. 2.2.3).
The overall effect of enhancing recovery through an increase in
the microscopic displacement factor is to shift the range of values
0.5 < Eps < 0.75 to 0.7 < Eps < 1.
1.0 100
IFT(mN/m)
mo/mw = 1.5 s = 0.1
mo/mw = 1.38 s = 0.3
Water Relative Permeability (Krw)
0.8 80
Oil Relative Permeability (Kro)
mo/mw = 1.28 s = 34
Oil Recovery, Percent. PV
0.6 60
0.4 40
Legend
0.2 20 s = 34 mN/m
s = 0.03 mN/m
s = 0.01 mN/m
0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 0 1 2 3 4 5
Water Saturation, SW Pore Volumes Injected, Vpi
L
x=
2 w
xf fluid Flo
x= d
p lace
Dis
1
uid
0 ing fl
x= lac
p
Dis ε
α = dip angle
to remember that the exact pattern of viscous fingers that will form
in a given displacement is impossible to predict (they are random)
although it is possible to predict their average behaviour (e.g. Todd
and Longstaff, 1972; Koval, 1963; Homsy, 1987). This contrasts with
channelling caused by rock heterogeneity in which the displacing
fluid always follows the same path through the rock, the path being
determined by the permeability pattern.
Viscous fingers are most likely to form in miscible displacements
although they can also form in immiscible displacements, provided
that the viscosity ratio between the displacing fluid and the oil
is sufficiently high. A formal derivation of this condition can be
obtained by considering small perturbations to the location of a
displacement front, Fig. 2.2.4. In particular, one derives expressions
for the movement of the bulk front, dxf /dt, and compares that with
the movement of the perturbation, d(xf + )/dt. The condition for
instability is given by,
d d(xf + ) dxf
= − > 0. (2.2.4)
dt dt dt
We do not have time to cover the derivation in this module,
but a remarkably simple condition falls out of this analysis. In
miscible displacements, ignoring gravity, the condition for stability
is dependent on the viscosity ratio
µo
M= ≤ 1. (2.2.5)
µd
To determine whether viscous fingers are likely to form (where µo is
the oil viscosity and µd is the viscosity of the miscible displacing
fluid). In immiscible displacements the condition for stability is
dependent on the shock front mobility ratio,
kr,w (Swf ) k (S )
µw + r,oµo wf
Msf = kr,o (Swc )
≤ 1, (2.2.6)
µo
where krw (Swf ) and kro (Swf ) are the water and oil relative perme-
abilities at the shock front saturation, Swf , and kro (Swc ) is the oil
relative permeability at the connate water saturation.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 61
Viscous fingers will form if M or Msf > 1 and the higher the
mobility ratio, the lower the sweep (Fig. 2.2.5). It is important to
remember that a 1D Buckley–Leverett analysis does not capture
the effect of macroscopic viscous fingering even if there is a large
viscosity ratio between the oil and the displacing fluid as viscous
fingering is a 3D phenomenon. It should also be noted that it is
very hard to explicitly model viscous fingers using conventional
reservoir simulators. You need a very fine grid and random perme-
ability or saturation fluctuation to trigger the fingers (Fig. 2.2.6).
It is more usual to capture the average effects of viscous fingers
using an empirical model such as the (Todd and Longstaff, 1972)
model.
80 38.2
Area Contacted by Drive, % Total Area
71.5
70
60
50
40
30
20
10
Figure 2.2.5. The impact of viscous instability on sweep for a miscible displace-
ment in a quarter five spot pattern.
Source: Haberman (1960).
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 62
M= M= M=
PVI = . PVI = . PVI = .
Oil saturation
Figure 2.2.6. Viscous fingering reduces the aerial sweep in a Hele Shaw cell.
Source: Djabbarov (2014).
z Gas s
q
α
h h
+ ∆h
∆ Oil
ht
l
Θ
Leng
th
Leng
th
Leng
th
et al., 1990; Fayers and Muggeridge, 1990; Tchelepi and Orr, 1994).
This is particularly important in gas injection processes and CO2
sequestration when the injected gas migrates quickly to the top of
the reservoir and then flows rapidly along the top, forming a gravity
tongue (Fig. 2.2.7). This effect may be less in a dipping reservoir
if the gas is injected at the top of the dipping layer and if the
vertical permeability is low. It is possible to estimate whether a
gravity tongue will form using the dimensionless number (Fayers and
Muggeridge, 1990).
u(1 − 1/M )µo h
Rv/g = 2 −θ , (2.2.8)
∆ρgkz L
where ∆ρ is the density difference between the oil and the gas (kg
m−3 ), kz is the vertical permeability (m2 ), h is the reservoir thickness
(m), L is the well spacing (m) and θ is the dip (radians). Note, this
can only reliably calculate the effects of dip when θ < 0.1 (about 10◦ ).
The displacement will always be stable if Rv/g < 0 (because M < 1
or because of a large dip). A gravity tongue will form if 0 < Rv/g <
1, for 1 < Rv/g < 10 flow will be influenced by both gravity and
viscous effects whilst the displacement will be viscous dominated if
Rv/g > 10. It should be noted that it is very difficult to use numerical
simulations to model unstable gravity tongues because a high degree
of vertical refinement is needed to capture the leading edge of the
tongue, instead it may be better to use a vertical equilibrium option
(if available in your simulator) if the above calculation suggests flow
will be gravity dominated. As with viscous instability the main way to
manage gravity segregation is through control of the mobility of the
injected fluid, typically this is via water alternating gas injection. The
water tends to reduce the mobility of the gas and, even if it does not
fully prevent the formation of a gravity tongue, it will improve sweep
as the water will tend to sink towards the bottom of the reservoir
whilst the gas rises to the top.
Geological heterogeneity is perhaps the most usual cause of
reduced macroscopic sweep although it can, on occasion, improve
sweep e.g. a fining upwards sequence will reduce the tendency of
gas to form a gravity tongue above oil during gas injection. EOR
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 65
Vdp = 1 − e−Sk ,
whereas,
1
n
1
Sk = 1+ (ln ki − ln k),
4(n − 1) n−1
i=1
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 66
C
.
.
. .
.
. .
. . .
. .
C C
Phase 1: C rich
c , c ,c
a Plait point
c ,c ,c
Single-phase
Phase 2: C rich
C C
Heavy pseudo component, Crude oil composition Intermediate pseudo
e.g. C7+ component, e.g. C6–
Figure 2.3.3. A two-phase ternary diagram. A fluid mixture with overall com-
position a will exist as a mixture of two phases with component concentrations
given by the ends of the tie line on the binodal curve.
Mixing
line
Resident
oil
C C
Figure 2.3.4. A two-phase ternary diagram showing the mixing line for the
injection of C1 , first contact miscible with the resident oil.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 74
Plait Point
Plait Point
l
ica
rit
LL fC
O e
n
io Lin
e ns Tie
Ext
100 100
Crude
Vol % Vol %
C+2 C–6
100 Vol % CO2
Plait
Point
Li f
Ti n O
ne
al o
ic si
e
rit ten
Ex
100 100
C
Crude
Vol % Vol %
C+2 C–6
Figure 2.3.5. Ternary diagrams of mixtures of CO2 with Wasson crude at 1,350
psi (above) and 2,000 psi (below). Both are at 105◦ F. Increasing pressure results
in a larger field of potential miscibility with the injected fluid.
Va mpos
co
po itio
uri
sin n
B
gg
as
A D
Plait
C
E point
Miscible
Resident
oil
C C
Figure 2.3.6. A two-phase ternary diagram showing the development of miscibility
through a vapourising gas drive, injecting C1 , multi-contact miscible with the
resident oil.
Injected
gas
A C E
B
D
Miscible
Resident Condensing oil
oil composition
C C
on opposite sides of the critical tie line. This is an imaginary tie line
that goes through the Plait point but is parallel to the tie lines in
the two-phase region just before the Plait point is reached.
Through this simple graphical analysis of a gas injection process
it is thus possible to tell a lot about the system, whether there are
opportunities for a vapourising or condensing gas drive and if the
composition of an injected gas could be tailored, by diluting for
example with some amount of the component C2 , so that miscibility
develops. At the same time it is important to keep in mind the
limitations of such a simplistic analysis. Ternary diagrams only apply
at a single pressure and temperature and we have not discussed
at all the significant efforts required through modelling and in the
laboratory to obtain such diagrams, particularly where crude oil is
concerned. The multiple contact process is not a discrete process as
was shown here, rather the gas-rich or oil-rich phases are continuously
evolving as they contact more resident oil or injected gas. Fig. 2.3.9
shows the composition path for a displacement of crude oil by CO2
simulated in a 1D displacement scenario (Gardner et al., 1981). Our
analysis was 0D and did not consider how fluids flow and contact
each other, nor any time constraints on the processes of chemical
71 Vol % UL 73 Vol % UL
12.2 Vol %
Li f
Ti n o
ne
al io
ic s
e
45.5
rit en
Vol % 35.5 LL
C Ext
‘High’ Dispersion
Level Low (Experimental)
Dispersion Level
Plait Point
Figure 2.3.9. A two-phase ternary diagram from Gardner et al. (1981) showing
the simulated composition path in a 1D displacement for the process shown in
Fig. 2.3.8.
Plait
point
Critical
tie line
r
r
C C
∂Si ∂fi
+ = 0, (2.3.5)
∂tD ∂xD
λi
fi = ,
i λi
kr (Si )
λi = .
µi
dfi
xD = (Si )tD , (2.3.7)
dSi
which says that a saturation, Si , will propagate forward in the
medium at a rate given by the derivative of the fractional flow
function, dfi /dSi , evaluated at that saturation. The derivative of
the fractional flow function is thus referred to as the dimensionless
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 82
∗ df
f1 − f1o
vD = S =S ∗ = , (2.3.8)
dS1
1 S1 − S1o
S1 =S ∗
where S1o and f1o are the saturation and corresponding fractional
flow of phase 1 at an initial condition prior to displacement. In a
graphical representation, S ∗ is the saturation at which the slope of
the fractional flow curve is equal to the slope of the straight line
passing through the points (S1o , f1o ) and [S1∗ , f1 (S1∗ )].
∂Ci
=φ. (2.3.12)
∂t
This must be balanced by fluxes of the component into and out of
the volume. In this class we will consider a 1D system where the net
flux is given by,
∂
ci,j qj (2.3.13)
∂x
j
∂
= ci,j fj qt (2.3.14)
∂x
j
∂Fi
= qt , (2.3.15)
∂x
for incompressible fluid systems where ∂q t
∂x = 0. Note that volume is
often not conserved when components transfer phases, i.e. the density
of CO2 in a hydrocarbon rich phase may well be very different than
the density of CO2 in the CO2 rich phase, and so this is not always
appropriate.
The conservation equation is thus,
∂Ci ∂Fi
+ qt
φ = 0. (2.3.16)
∂t ∂x
We can convert this to dimensionless units using the chain rule,
∂Ci ∂Ci ∂tD qt ∂Ci
= = , (2.3.17)
∂t ∂tD ∂t φL ∂tD
∂Fi ∂Fi ∂xD 1 ∂Fi
= = . (2.3.18)
∂x ∂xD ∂x L ∂xD
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 84
fw
.
water–oil
. water–solvent
. . . .
Sw
Figure 2.3.11. Fractional flow curves for an oil phase displacing water where the
water–solvent curve has a higher mobility ratio.
Figure 2.3.12. Showing the solvent and water injection velocities at an injection
fractional flow where the solvent velocity will be larger than the injected water
velocity
Figure 2.3.13. Showing the solvent and water injection velocities at an injection
fractional flow where the solvent velocity will be less than the injected water
velocity.
Figure 2.3.14. Showing the solvent and water injection velocities at an injection
fractional flow where the solvent velocity is equal to the injected water velocity.
So = 1 − Sw − Sg . (2.3.30)
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 91
in (Blunt, 2000; Juanes and Patzek, 2004; Baker, 1988; Delshad and
Pope, 1989). This is an area of ongoing research but in summary,
models currently considered the most accurate for representing three-
phase relative permeability interpolate between values of two-phase
relative permeabilities. In other words, for the oil relative perme-
ability an interpolation is made between the relative permeability in
the oil–water system, kr,o(w) and in the oil–gas system, kr,o(g) , with
the interpolation weighted by the water and gas saturations. The
equivalent can be done for the water and gas relative permeabilities
(Blunt, 2000). Following the model proposed by (Baker, 1988), the
relative permeabilities for oil, water and gas can be given as
(Sw − Sw,i )kr,o(w) + (Sg − Sg,r )kr,o(g)
kr,o = , (2.3.38)
(Sw − Sw,i ) + (Sg − Sg,r )
(So − So,i )kr,w(o) + (Sg − Sg,r )kr,w(g)
kr,w = , (2.3.39)
(So − So,i ) + (Sg − Sg,r )
(Sw − Sw,i )kr,g(w) + (So − So,i )kr,g(o)
kr,g = . (2.3.40)
(Sw − Sw,i ) + (So − So,i )
In this model, Sw,i is the irreducible water saturation, Sg,r is the
residual gas saturation in an oil–water displacement and So,i is the
initial oil saturation in a gas–water displacement, the latter two are
usually zero. Using this model, the three-phase relative permeability
functions are obtained through the measurement of six two-phase
relative permeability curves, both phases in an oil–water, oil–gas and
gas–water displacement.
flooding although it has only been used in field pilots to date. It was
devised to combine the benefits of using a surfactant to reduce the
oil–water interfacial tension and hence reduce capillary trapping of
the oil and a polymer to provide mobility control whilst including an
alkali to prevent both of these being lost by adsorption onto the rock.
In contrast, low salinity water injection has only been extensively
discussed since it was rediscovered by (Yildiz and Morrow, 1996).
There have been several single well field tests as well as one interwell
field trial (Seccombe et al., 2010) but it will not be deliberately
deployed until Clair Ridge comes on stream in 2017. High sulphate
water injection is a specialised EOR technique for chalk reservoirs
e.g. (Zhang et al., 2007). Only polymer flooding and low salinity
water flooding will be discussed in detail in this section.
above with two key differences: first, the dissolved chemical con-
centration, C1 , of interest is polymer and it is dissolved in the
aqueous phase. Second, one of the key processes to capture in
polymer flooding is the reversible adsorption of the polymer onto
solid walls of the rock pores. This is done by accounting for some
amount of the concentration, C1 , as made up of the adsorbed
concentration, c1a , the amount of adsorbed polymer per unit pore
volume. Figure 2.4.1 shows two fractional flow curves representing
the displacement processes between the water–oil system and the
aqueous polymer solution–oil system, with the mobility lower in
the polymer solution displacement due to the increased viscosity.
In this iteration thus the total concentration, including the adsorbed
component, is given by,
c1a = Rc1w ,
fw dfw
= .
Sw,3 + R dSw
Sw =Sw,3
Figure 2.4.2. Fractional flow analysis for low salinity water flooding applied for
primary recovery.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 99
of this line with the high salinity fractional flow curve (Sw,2 , fw,2 )
gives the shock front saturation of the leading edge of the connate
water bank (Sw,2 ) and the point of tangency gives the shock front
saturation of the leading edge of the low salinity water (Sw,3 ). The
characteristic velocity of the leading shock is determined by drawing
a line from the connate water saturation of the high salinity fractional
flow to the point (Sw,2 , fw,2 ) and calculating the gradient of that
line. The characteristic velocity of the trailing shock is simply the
gradient of the original tangent line. Finally, there is a rarefaction
to the residual oil saturation. If low salinity flooding is applied as
a secondary or tertiary recovery process, the oil saturation will be
low. This will result in the creation of an oil bank, characteristic of
enhanced recovery induced by a wetting state change (Fig. 2.4.3).
Jerauld et al. (2008) proposed a refinement of this model when
implementing it in a numerical simulator. They asserted that the
injected low salinity water would only begin to alter oil recovery
when the water salinity was lower than a specified upper salinity
threshold (e.g. 7,000 ppm) and that it would only be fully effective
once salinity was below a lower threshold (e.g. 1,000 ppm). For
salinities below the lower threshold then the flow would be described
by the more water-wet relative permeabilities whilst for salinities
above the upper threshold the flow would be described by the more
oil-wet relative permeabilities. For intermediate salinities then the
relative permeabilities would be a linear combination of the high
Figure 2.4.3. Fractional flow analysis for low salinity flooding applied as a
secondary or tertiary recovery mechanism.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 100
oil in pores and bypass oil in lower permeability zones. EOR injectant
applied after a waterflood may then simply follow the flow paths
already established by the injected water so less oil will be accessed.
It will also take longer for the oil that is swept up by the EOR
injectant to reach the production well. Nonetheless, many engineers
will only consider using EOR after a waterflood. This may be because
at the time the field was discovered the oil price was too low to
justify the additional expense of an EOR scheme. However, even
when the oil price has been higher many engineers have preferred
to start production using water injection. In this case, they reason
that this will give them a better understanding of the geological
heterogeneity in their reservoir and will thus reduce the risk of a
poor outcome when they finally deploy EOR. It is not always clear
that this reduction in perceived risk outweighs the loss of additional
recovery from applying the EOR scheme in secondary mode. The
response from tertiary EOR schemes may be so slow that it can take
a year or more for there to be a noticeable slowing of the decline in oil
rate or indeed an observable improvement in oil rate. As an example
it took more than two years from the start of gas injection in Magnus
to see a noticeable improvement in oil recovery (Brodie et al., 2012).
It can also be difficult to clearly relate this improvement in oil rate
to the EOR process rather than more recent activities such as well
workovers. In any case, this delay may make the scheme uneconomic.
If it is likely that EOR will be applied post-waterflood then it
is important to plan the field infrastructure to allow for this. If this
is not done then it may become prohibitively expensive to retrofit
these facilities e.g. many offshore platforms on older fields in the
North Sea have no space for the installation of desalination plants
or the preparation of polymer solutions. Injection of carbon dioxide,
whilst potentially effective in terms of recovering additional oil and
also providing a means of mitigating climate change through the
storage of the gas, would require installation of specialised pipework
and wells that are corrosion resistant. To give an indication of the
potential costs, in 2014 drilling a new well in the North Sea could
cost £100 million. In the case of CO2 EOR/sequestration there
would also be the additional cost of a subsea pipeline network to
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 103
distribute the CO2 to fields from the power stations where it would be
captured.
Leaving aside economic considerations, the appraisal, planning
and development of EOR schemes can be significantly more complex
than for a conventional water flood. We have already seen in the
previous chapters how 1D fractional flow analysis for WAG, polymer
and low salinity water injection is more involved than for water
flooding, not to mention the additional understanding of phase
behaviour that is needed when planning a miscible gas injection
scheme. Leaving aside the logistical considerations discussed in
preceding paragraphs and the additional facilities design required,
further specialised laboratory tests may be needed to determine
EOR specific data such as the parameters needed to describe phase
behaviour when injecting a miscible or multi-contact miscible gas,
the relative permeability curves needed to plan a low salinity flood
in the reservoir of interest or to quantify the expected levels of
polymer adsorption. Having estimated the microscopic displacement
efficiency of possible EOR schemes using the analytical techniques
described in previous chapters, it will then be necessary to assess
the impact of geological heterogeneity on macroscopic sweep as well
as the range of possible outcomes due to the uncertainty in that
heterogeneity. This requires the use of specialised reservoir simulation
software which in turn relies on specialised reservoir engineering
expertise (e.g. chemistry of rock–fluid interactions for low salinity
waterflooding, chemistry and flow behaviour of the emulsions that
form during ASP flooding, phase behaviour of multi-contact miscible
displacements and use of equations of state). In some cases this
additional complexity may be sufficient to dissuade engineers and
their managers from even considering an EOR scheme.
Despite these various economic, practical and technical diffi-
culties EOR still has the potential to significantly improve oil
recovery. McCormack et al. (2014) suggest that EOR could recover
an additional 0.6–1.2 billion barrels of oil, from the UKCS alone.
Worldwide it could increase the reserves to production ratio from 40
years to perhaps 45 or 50 years. This can only be achieved, however,
through a combination of good reservoir engineering, longer term
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 104
References
Al-Saadi, F. S., Amri, B. A., Nofli, S., Wunnik, J. V., Jaspers, H. F., Harthi,
S., Shuaili, K., Cherukupalli, P. K. and Chakravarthi, R. (2012). Polymer
flooding in a large field in South Oman — initial results and future plans.
SPE EOR Conference at Oil and Gas West Asia held in Muscat, Oman,
(SPE 154665).
Aleklett, K., Hook, M., Jakobsson, K., Lardelli, M., Snowden, S. and
Soderbergh, B. (2010). The peak of the oil age — analyzing the world
oil production reference scenario in World Energy Outlook 2008. Energ.
Policy, 38(3), 1398–1414.
Anderson, W. (1987). Wettability literature survey — part 6: The effects of
wettability on waterflooding. J. Petrol. Technol., 39(12), 1605–1622.
ARI (2009). CO2 storage in depleted oilfields: Global application criteria for
carbon dioxide enhanced oil recovery. Technical Report 2009–12, IEA
Greenhouse Gas Program.
Baker, L. E. (1988). Three phase relative permeability correlations. SPE/DOE
Enhanced Oil Recovery Symposium Tulsa, Oklahoma, (SPE/DOE 17369).
Blunt, M. J. (2000). An empirical model for three-phase relative permeability.
Soc. Petrol. Eng. J., 5(4), 435–445.
Blunt, M. J. (2013). Reservoir Performance Predictors Course Notes. Imperial
College London.
Brodie, J., Jhaveri, B., Moulds, T. and Mellemstrand, H. S. (2012). Review of
gas injection projects in BP. In Proceedings of the 18th SPE Improved Oil
Recovery Symposium, Tulsa, OK 14–18 April.
Chang, H., Zhang, Z., Wang, Q., Zu, Z., Guo, Z., Sun, H., Cao, X. and Qiao,
Q. (2006). Advances in polymer flooding and alkaline/surfactant/polymer
processes as developed and applied in the People’s Republic of China.
J. Petrol. Technol., 58(2), 84–91.
Christie, M., Jones, A. and Muggeridge, A. (1990). Comparison between labora-
tory experiments and detailed simulations of unstable miscible displacement
influenced by gravity. North Sea Oil and Gas Reservoirs-II.
Dang, C., Nghiem, L. X., Chen, Z. and Nguyen, Q. P. (2013). Modeling
low salinity waterflooding: Ion exchange, geochemistry and wettability
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch02 page 105
Smalley, P. C., Ross, B., Brown, C. E., Moulds, T. P. and Smith, M. J. (2009).
Reservoir technical limits: A framework for maximizing recovery from oil
fields. SPE Reserv. Eval. Eng., 12(04), 610–617.
Sorbie, K. S. (1991). Polymer-Improved Oil Recovery. Springer.
Sorbie, K. S. and Seright, R. S. (1992). Gel placement in heterogeneous systems
with crossflow. Proceedings SPE/DOE Enhanced Oil Recovery Symposium,
Tulsa, OK, (SPE-24192-MS).
Stoll, W., al Shureqi, H., Finol, J., Al-Harthy, S., Oyemade, S., de Kruijif, A.,
van Wunnik, J., Arkesteijn, F., Bouwmeester, R. and Faber, M. (2011).
Alkaline/surfactant/polymer flood: From the laboratory to the field. SPE
Reserv. Eval. Eng., 14(06), 702–712.
Taber, J., Martin, F. and Seright, R. (1997a). Eor screening criteria revisited —
part 1: Introduction to screening criteria and enhanced recovery field
projects. SPE Reservoir Eng., 12(3), 189–198.
Taber, J., Martin, F. and Seright, R. (1997b). EOR screening criteria revisited —
part 2: Applications and impact of oil prices. SPE Reservoir Eng., 12(03),
199–205.
Tang, G. Q. and Morrow, N. R. (1999). Salinity, temperature, oil composition
and oil recovery by waterflooding. SPE Reservoir Eng., 12, 269–276.
Tchelepi, H. A. and Orr, F.M. Jr. (1994). Interaction of viscous fingering,
permeability heterogeneity, and gravity segregation in three dimensions.
SPE Reservoir Eng., 9(04), 266–271.
Thyne, G. (2011). Evaluation of the effect of low salinity waterflooding for 26 fields
in Wyoming. SPE Annual Technical Conference and Exhibition, Denver,
Colorado, (SPE 147410).
Todd, M. and Longstaff, W. (1972). The development, testing and application of
a numerical simulator for predicting miscible flood performance. J. Petrol.
Technol., 24(7), 874–882.
USEIA (2014). Annual Energy Outlook 2014 with projections to 2014. Technical
Report DOE/EIA-0383(2014), US Energy Information Administration.
Wallace, M. and Kuuskraa, V. (2014). Near-term projections of CO2 utilization
for enhanced oil recovery. Technical Report DOE/NETL-2014/1648,
National Energy Technology Laboratory.
Webb, K. J., Black, C. J. J. and Al-Ajeel, H. (2003). Low salinity oil recovery- log
inject log. Middle East Oil Show, Bahrain, 9–12 June, (SPE-81460-MS).
Yildiz, H. O. and Morrow, N. R. (1996). Effect of brine composition on recovery
of Moutray crude oil by waterflooding. J. Petrol. Sci. Eng., 14, 159–168.
Zhang, P. M., Tweheyo, M. T. and Austad, T. (2007). Wettability alteration and
improved oil recovery by spontaneous imbibition of seawater into chalk:
impact of the potential determining ions Ca2+ , Mg2+ and SO−4 2 . Colloid
Surface A, 301, 199–208.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 109
Chapter 3
Numerical Simulation
Dave Waldren
Petroleum Consulting and Training Ltd.
Cobham, Surrey, UK
109
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 110
MASS BALANCE
x x + dx
x u
x x x y z t x ux y z t
x y z( t t t )
( u) ( )
x t
k
u ( p gz )
l
uw ∝ (h1 − h2 ). (3.2.3)
k ∂
u= (p + ρgz). (3.2.4)
µ ∂l
α v3
mgSinθ
kko ∂
∇ ∇(po − ρo gz) − qo = (ρo φSo ), (3.2.12)
µo Bo ∂t
kkg kko
∇ ∇(pg − ρg gz ) + ∇ Rs (p)∇(po − ρo gz) − qg
µg Bg µo Bo
∂
= {(Rs ρo So + ρg Sg )φ}, (3.2.13)
∂t
kkw ∂
∇ ∇(pw − ρw gz) − qw = (φρw Sw ), (3.2.14)
µw Bw ∂t
So + Sw + Sg = 1.0, (3.2.15)
These six equations and six unknowns represent the model for fluid
flow in the reservoir but, due to coupling (Rs depends on p, pw
depends on po as well as pcow (Sw) and the nonlinearity, (ko depends
on So , Bo depends on po ) in general no analytical solution is possible.
However an approximate solution can be obtained for these
equations at preselected space points and for discrete time values
by using the method of finite differences.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 119
ui+1 − ui = δu . (3.2.18)
∂f f f (ui+1 ) − f (ui )
= + O(δu), (3.2.21)
∂u δu
∂f b f (ui ) − f (ui−1 )
= + O(δu), (3.2.22)
∂u δu
∂2f f (ui−1 ) + f (ui+1 ) − 2f (ui )
= + O((δu)2 ), (3.2.23)
∂u2 δu2
i+1
i
f (u)
i 1
where for each space point we have a similar difference equation for
each of the three fluid components.
Multiplying through by the volume of the element converts the
specific rate term to mass per unit time and gives:
δyδz kko kko
(pi+1 − pi ) − (pi − pi−1 )
δx µo Bo
i+1/2 µo Bo i−1/2
V φSo n+1 φSo n
−qo = − . (3.2.25)
δt Bo i Bo i
EXPLICIT FORMULATION
Ak ∆p 1
δρ = δt ,
µo Bo ∆x A∆xφ
∆x2 φµo c
δt ≺ . (3.2.26)
2 k
Ak ∆p
N= δt (3.2.28)
µo Bo ∆x
Ak ∆p 1
δp = δt . (3.2.29)
µo Bo ∆x A∆xφ
k ∆p
δp = δt. (3.2.30)
φµo c ∆x2
∆x2 φµo c
δt < . (3.2.31)
2 k
Using:
∆x = 104 cm
µo = 1cp
c = 10−4 atm−1
k = 1darcy
Φ = 0.2
Unstable
Explicit
q12
Stable
Explicit
Stable Implicit
Time 1 Day
k
ai− = δtδyδz, (3.2.34)
µo Bo i−1/2
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 124
k
ai+ = δtδyδz, (3.2.35)
µo Bo i+1/2
1
1 ∂φ ∂ Bo
ai = a− + a+ + −φ V δx. (3.2.36)
Bo ∂p ∂p
∆pi = pn+1
i − pni . (3.2.38)
A∆p = b, (3.2.40)
Theory
Theory
Wcut
Sw
Simulation Simulation
x Time
A∆p = b (3.2.41)
k
1 1 1 / Bo
pk 1
Bok 1
Bo ( p k ) p
stops when
• all the values of δp are smaller than some tolerance, and
• the material balance is within a tolerance, or
• the maximum number of iterations has been reached.
The iterative procedure is illustrated in Fig. 3.2.8.
If, for some cells, the value of the formation volume factor at the
new time level, B n+1
o , is not convergent, then the densities used in
the difference equation will be incompatible and the solution will not
conserve mass. This will result in a material balance error.
1 2 3 4 5
6 7 8 9 10
11 12 13 14 15
1 4 7 10 13
2 5 8 11 14
3 6 9 12 15
or
U ∆p = L−1 b (3.2.46)
∆p = U −1 L−1 b. (3.2.47)
w = 2ny nz + 1. (3.2.48)
N ∝ niter nx ny nz , (3.2.50)
Compute matrix
elements
Set up linearised
matrix equation
Outer
Perform 1 Gauss-Seidel Iterations
pass
Inner
No Iteration
Check for convergence
of linear solver
Yes
No
Check for convergence
of
of non-linear
nonlinear effects
effects
Yes
Output solution
Fw
Sw Sw shock
Sw shock
Sw
3.4.1.2. Initialisation
In order to set the initial phase saturations and pressures, the model
requires
3.4.1.5. Output
The user can request different types of reports or data to be output
• time step summary,
• well summary,
• region summary,
• cell by cell values of variables,
• graphics files,
• pseudo files,
• restart files.
The detailed form of the input data depends on the type and
input format specification of the model which is being used. As the
complexity of the physical processes treated by the model increases
so does the quantity of data require to define the processes. Although
most simulators provide some sort of data screening, they are by no
means infallible and the requirement for consistency in the input data
is the responsibility of the engineer setting up the model.
equation becomes
1 ∂ ∂p φµc ∂p
r = . (3.5.2)
r ∂r ∂r k ∂t
This may be discretised in the same way as the Cartesian equation,
giving
1 kko
r (pi+1 − pi )
ri δri µo Bo i+1/2 i+1/2
kko
− r (pi − pi−1 )
µo Bo i−1/2 i−1/2
V φSo n+1 φSo i
= − , (3.5.3)
δt Bo i Bo n
Clearly the mesh size should also depend on the amount and quality
of data available. There is little point in constructing a detailed
model with many thousands of grid cells for a prospect with only
the discovery well and coarsely spaced 2D seismic data available.
Cross flow
Pinch out
Fault
This can have severe consequences for the speed of each inner
iteration and also on the convergence properties of the matrix.
Oil zone
Transition Zone
Well
Aquifer
1 2 3 4 5
6 7 8 9 10
15 16
11 12 13 14 17 18
Figure 3.5.7. Local grid refinement showing the effect on the matrix structure.
Gathering centre 1
3.5.9. Non-Orthogonality
In Sec. 3.5.3, we described the use of stream lines for setting up
the spatial discretisation. This system is useful only in so far as the
flow is along the stream lines and such a grid system would not
be chosen for a different application. This is because the choice of a
non-orthogonal system should lead to transmissibility terms between
cells not usually connected in the normal five point differencing
schemes. However, for the stream line choice, the flux in these
directions is zero so that neglecting these terms does not lead to
errors.
For an application with reasonably fixed flow patterns (for
instance the simulation of a symmetry element of a pattern develop-
ment) this procedure can be very useful but care must be taken that
field operating conditions do not seriously alter the stream lines.
Although distorted coordinate systems may give an improved
geometrical representation of the reservoir and its fault system, it
may also introduce a further level of divergence from the differential
equation due to neglecting some cross-term transmissibilities. In this
respect it is similar to an additional space truncation error.
The same is true of the case where local grid refinement is used
as the fluxes between the fine grid area and the coarse grid area are
subject to additional truncation error terms which depend on the
degree of refinement.
However, apart from trying different grid systems (which is not
frequently done), there is no a priori way of estimating the size of
the error introduced.
A final comment on orthogonality of reservoir models is that the
choice of axes, x along the dip, y along the strike and z vertical
is non-orthogonal, except for horizontal models, and that the error
introduced is usually small compared to all the other approximations
used and the accuracy of the data available.
3.6.1.1.1. Porosity
The porosity of interest to reservoir engineers is the value of
the connected porosity (which may be different from the absolute
porosity as measured by logs). This is rapidly measured by cleaning
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 149
and drying the core sample and comparing the weight of a helium
container with and without the core. This gives a value for the grain
volume and, knowing the value for the bulk volume, the connected
porosity can be calculated.
An additional estimate of porosity is often available from the
amount of fluids which were recovered from the core sample.
3.6.1.1.2. Permeability
The cleaned and dried core is inserted into a sleeve and a gas (usually
air) is passed through the core. The pressure difference is adjusted
to give a measurable flow rate from which the permeability can be
calculated. The dependence of rate on permeability for a low pressure
gas (density inversely proportional to pressure) becomes
kA
q= (p2 − p22 ), (3.6.1)
2pbase Lµ 1
where pbase is the base pressure and L is the length of the core plug.
The viscous flow of gases and liquids is not quite the same due to
the different scale of the boundary layers. Klinkenberg showed that
the gas permeability was related to the liquid permeability by
c
kg = kl 1.0 + , (3.6.2)
p
3.6.1.3.1. Compressibility
In situ, rock is subjected to forces from the over burden and from the
fluid occupying the pore space. The difference between these is the
net confining pressure. As the pore pressure decreases, allowing rock
grains to expand, the net confining pressure increases, tending to
reduce the bulk volume of the rock. Both of these effects reduce the
porosity and this effect is included in the formulation of the finite
difference equations, usually as an elastic process with a constant
compressibility.
Experimentally, the core sample is subjected to triaxial loading
such that there is no deformation in the lateral direction and the
change in volume is recorded as a function of the net confining
pressure.
The values obtained should be used to compute the derivative of
porosity with respect to pressure at the initial reservoir net confining
pressure and corrected to give the uniaxial result by using the
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 151
Recompression
∆θ
Unloading
θ Burial
Grain pressure
correction factor:
1 1+v
fcorr = ctriaxial , (3.6.3)
3 1−v
where v is Poisson’s ratio for the rock (approximately 0.3).
This gives a correction factor fcorr of approximately 0.6.
The effect of net confining pressure should be included in the
values of porosity and (sometimes) permeability which are input to
the model to correspond to the initial conditions.
If the reservoir to be simulated has a component of compaction
drive then the effective rock compressibility can be an order of mag-
nitude greater than the usual value (3 · 10−6 psi−1 to 7 · 10−6 psi−1 ).
The effect of compaction is not reversible so the standard elastic
treatment is not sufficient and a special option dealing with this
phenomenon should be used.
The effects of inelastic behaviour are demonstrated in Fig. 3.6.1.
Kro(Swc)
Kro Krw
Krw(1-Sorw)
Swc 1-Sorw
1.0
1.0
Krg(Swc)
Kro(Swc)
Krg
Kro
Kro(Sgc)
Krg(Sorg)
Sorg Sgc
Swc
at laboratory conditions and the fluids used are water (or simulated
formation water) and a standard mineral oil with a viscosity of
approximately 20 centipoise.
This choice of viscosities permits the initial conditions of connate
water saturation to be recreated in a short period of time and also
produces rapid water breakthrough so that the full range of the
relative permeability function can be measured.
This experiment is performed at high velocity in order to
minimise the effects of the capillary pressure discontinuity at the
entrance and exit faces of the core.
A more time consuming process is to measure the steady state
relative permeability by simultaneous injection of water and oil into
a core at rates corresponding to the velocity of fluids in the reservoir.
In order to remove the capillary end effects, long cores can be used.
This experiment may be carried out at laboratory conditions or at
reservoir temperature and pressure with live crude and simulated
formation water.
Because there is no longer a Buckley–Leverett displacement, the
relative permeability function can be measured for the equilibrium
core saturation which is set up as a result of the relative
permeabilities.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 154
pc = po − pw , (3.6.5)
Sand Grain
Oil
r1
Water
r2
(a) (b)
Pc
Drainage
Imbibition
0 Swc Sw 1.0
combinations normally used are an air brine system, where air is the
non-wetting phase, or a mercury air system where air is the wetting
phase.
In the air brine experiment, air is forced into the core reducing
the initially high water saturation as the pressure increases. In the
mercury air experiment, mercury is forced into the core plug and the
saturation of mercury versus the mercury pressure is recorded.
The mercury injection experiment is frequently used as it is
rapid to perform but it has the disadvantage that the core sample is
permanently damaged and is of no further use.
The neutron tool relies on the kinematic effect that the target
with the greatest effect is one with the same mass as the projectile.
This means that the neutron tool is sensitive to hydrogen atoms and
hence to hydrocarbons and water.
A combination of these tools then, indicating a reduction in
density and an increase in hydrogen content is indicative of a fluid
filled porous rock and the excursion of the curves is related to the
porosity and the fluid saturations. The response of the FDC–CNL
log for a porous interval is shown in Fig. 3.6.9.
The fluid saturations may be estimated from the resistivity
measurements because of the different electrical properties of drilling
mud filtrate, formation brine and hydrocarbons.
Measurement of the resistivity deep in the formation away from
the zone invaded by mud filtrate can give an estimate of the initial
saturations and the shallow resistivity can give an estimate of the
residual oil saturation. An example of the response to a water bearing
formation is shown in Fig. 3.6.9.
where αSw Vclay and βSw Vclay are the coefficients determined from the
log and core data cross-plots obtained from wells where correspond-
ing log and core data were available. An example of such a cross-plot
is shown in Fig. 3.6.10.
1000
100
Kair
10
1
10 14 18 22 25 30 34
Percent
Log (T+DT) / DT
∗ Sw − Swc
Sw = . (3.7.1)
1. − Swc − Sorw
The relative permeability values are then normalised according to
the end point values as follows:
∗)
kro (Sw
∗
kro (Sw )= (3.7.2)
kro (Swc )
and
krw (Sw∗)
∗
krw (Sw )= (3.7.3)
krw (1. − Sorw )
The curves resulting from this procedure for the oil relative perme-
ability are shown in Fig. 3.7.1.
Inspection of the resulting plots of normalised relative perme-
ability permits one or more representative or average curves to be
selected.
Different types of curve may be associated with say, different
values of net-to-gross ratio.
In order to use the curves as input to the model, the end point
values for the specific layer (or group of cells) must be specified.
These values can be estimated from inspection of cross-plots of the
relative permeability end points versus primary variables such as
permeability and porosity.
Usually a simple set of correlations (or values) suffices to give a
comprehensive set of relative permeability curves for the model e.g.
1000
100
Kair
10
1
0 20 40 60 80 100
Sw
1000
100
Kair
10
1
0 10 20 30 40 50 60
Sor
case for thermal simulation where there is close proximity of all three
phases.
the order in which they flood, then, when j layers have flooded, the
relative permeability values and average saturations are given by
j
i (1 − S i )
ki hi krw orw
i=1
k rw (S w ) = , (3.7.21)
j
ki hi
i=1
n
ki hi kro
i (S i )
wc
i=j+1
kro (S w ) = , (3.7.22)
n
ki hi
i=1
where
j
n
φi hi (1 − Sorw
i )+ φi hi Swc
i
i=1 i=j+1
Sw = . (3.7.23)
n
φi hi
i=1
T x1 K ro1 T x 2 K ro 2
K ro
Tx1 Tx 2
Kro Krw
Swc Sw 1-Sorw
3.7.6. Summary
The use of relative permeability arises because of the requirement
to take into consideration the consequences of saturation dependent
mobility at the microscopic scale.
These functions, when used on a macroscopic scale in reservoir
models embody the microscopic effects compounded with assump-
tions regarding the distribution of reservoir properties and the spatial
distribution of fluids within the finite difference cell.
The effects of the spatial distribution of fluids and the vertical
distribution of horizontal permeability can be accounted for by using
upscaling techniques which are representative of the large finite
difference cells.
The grouping of cells is made by ensuring that the up-scaled func-
tions used in the coarse model adequately represents the behaviour
of the high resolution model.
versus Sw
20
Average Curve
15
J (Sw)
10
0
0 20 40 60 80 100
Brine saturation percent
For any model layer with porosity and permeability φi and ki the
j function can be used to compute the capillary pressure as:
Sw − Swc
i φi
pic (Sw ) = j ∗ σres . (3.8.4)
1. − Swc
i ki
is different for every cell, this approach, although giving the correct
oil in place and stability at initialisation, is little used and has been
superseded by the slice integration technique.
3.8.3. Summary
The concept of capillary pressure, as with relative permeability, arises
from the microscopic effects of immiscible fluids occupying the same
pore space. When used in a finite difference model, the capillary
pressure. Along with the relative permeability, acquires a component
which describes the spatial distribution of fluids inside the finite
difference cell.
The user, by selecting different input for these functions can
impose diffuse flow, segregated flow or whatever he believes most
closely reproduces the reservoir behaviour, and to a large extent
it is these assumptions and the resulting pseudo functions which
determine the results of the model.
the boiling point line to a two phase region. This is shown for a
binary system of ethane and N -heptane in Fig. 3.9.1.
This two phase region also has limits in both temperature
and pressure, but unlike the single component where the limits in
temperature and pressure correspond to the critical point, the limits
for a mixture are not coincident with the critical point.
As may be seen in Fig. 3.9.1, the behaviour of a mixture of two
components is not a simple linear interpolation of properties and the
critical pressure for the mixture may be significantly greater than
that of either component.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 182
3.9.4. Definitions
Prior to discussing the experimental methods used to characterise
hydrocarbon systems, we will define some of the more important
terms used.
Saturation pressure
This is the pressure at which, during compression or expansion
at constant temperature, the gas system is in equilibrium with an
infinitesimal quantity of liquid or the liquid system is in equilibrium
with an infinitesimal quantity of gas.
P P
T T
P P
Volatile Condensate
T T
P P
T T
pV = znRT. (3.9.1)
The gas deviation factor and the gas formation volume factor are
related by
ps T
Bg = z, (3.9.2)
pTs
where ps and Ts are the pressure and temperature at standard
conditions.
Oil gravity
This is the ratio of the oil density to the density of water. This is
also quoted in terms of API gravity where:
141.5ρwater
γAPI = − 131.5. (3.9.3)
ρoil
Gas gravity
The gas gravity is the ratio of the density of the gas to the density of
air at standard conditions. At low pressure, the deviation factor for
both gases is approximately unity and so the gas gravity is given by
Mg
γg = (3.9.4)
Mair
where M is the molecular weight.
Isothermal compressibility
The isothermal compressibility is the relative change in volume per
unit pressure change:
1 dV
coil = − . (3.9.6)
V dp
The following experiments set out to determine values for these
quantities which specify to some reasonable accuracy, the behaviour
of the hydrocarbon system.
3.9.5. Experiments
For a black oil system, the experiments are designed to provide a
means of evaluating the pressure dependent fluid properties required
in the formulation of models from material balance to fully implicit
numerical models.
gas below the bubble point as the pressure decreases. This is shown
in Fig. 3.9.7.
In Sec. 3.9.2, we saw that the behaviour of a binary system
in terms of the component properties was not at all linear. By
performing the differential liberation experiment, at each pressure
step we create a new mixture and it would be unlikely that the
combined effect of the 10 pressure decrements would give the same
result as the constant composition expansion.
In fact, the total amount of gas released (both mass and volume)
tends to be larger for the differential liberation compared to the
constant composition expansion.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 190
Component
This is a constituent of the hydrocarbon mixture which might
correspond to a particular chemical species, a group of hydrocarbons
within a range of carbon numbers or a statistical mixture of com-
pounds evaluated by some means.
Phase
This refers to the state of a group of components which at the defined
temperature and pressure form a coherent mixture with measurable
properties. Thus in a saturated reservoir, the gas cap gas is an
example of a phase.
Phase equilibrium
This is the state of dynamic equilibrium achieved when at a
stable temperature and pressure, the compositions of the phases are
independent of time, that is all values of x and y are constant.
At equilibrium, we can define a relationship between the x and
y values for the various components as
yi = Ki xi , (3.10.1)
where
Ki = K(T, p, xi , xj , . . .) (3.10.2)
and K is called the equilibrium coefficient.
The computation of liquid vapour equilibrium is mathematically
difficult and in simulation of complex processes can consume a large
fraction of the computer time.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 192
Black oil
In a black oil model, it is assumed that the hydrocarbon system can
be adequately represented by two components with a distribution
of molecular weights. This distribution is that determined by the
separation facilities so that the two components correspond to stock
tank oil and gas (separator gas plus the gas evolved from the
separator liquid in going to tank condition).
The equilibrium coefficient for these two components at a fixed
reservoir temperature is assumed to be a function of pressure only.
In addition, all reservoir processes are assumed to be isothermal
and the temperature effects in producing to surface are accounted for
in the formation volume factor which relates surface to subsurface
volumes.
An assumption which is usually made (but can be relaxed) is
that the stock tank oil is non-volatile even at reservoir conditions
and that it does not partition into the vapour phase.
This implies that the vapour mole fraction for the gas component
is unity and that the reservoir gas has a fixed composition identical
to the surface gas.
Together with the assumption regarding the equilibrium
coefficient
we see that the ratio of the number of moles of stock tank oil in
the liquid phase to the number of moles of gas in the liquid phase is
determined uniquely by the equilibrium coefficient, K(p).
This means that for a given mixture of components at a set
pressure, knowing the form of K(p) (which is a transformation of
the solution gas content), the liquid vapour equilibrium mixture can
be computed a priori with no iterations.
Water is assumed not to partition into either hydrocarbon phases
and not to influence any of the hydrocarbon properties.
The data required for a black oil model is then
The formation volume factor is also scaled so that the bubble point
value corresponds to the flash experiment.
f
Bob
Bo (p) = Bod (p) d
. (3.10.6)
Bob
Here, the correct values have been taken as those from the constant
composition expansion (flash) experiment. If separator tests have
been made it is preferable to use those which most closely resemble
the field conditions.
After performing the scaling, we have corrected tables of data,
which for oil properties typically appear as in Figs. 3.10.1–3.10.3.
1 Bob − Bo (p)
co = , (3.10.7)
Bob p − pb
1 µo (p) − µob
cµo = . (3.10.8)
µob p − pb
For any new value of bubble point pressure, p∗b , then we can write
where
We = cumulative water influx,
U = aquifer influx constant,
WD = dimensionless water influx function,
tD = dimensionless time.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 199
The cumulative water influx, after n + 1 time steps, for a cell coupled
to the aquifer is
j=n
We (tn + 1) = U ∆pj WD (tDn+1 − tDj ). (3.11.2)
j=0
This requires that for each cell coupled to the aquifer, the history of
pressure changes must be stored and used to compute the current
value of the influx.
Because of the computational overhead involved, more approxi-
mate methods have been developed.
, (3.11.3)
pD (tDn+1 ) − tDn dp D
dt D n+1
where
∆pn+1 = total pressure drop,
pD (tD ) = dimensionless pressure.
The dimensionless pressure function is the constant terminal rate
solution to the radial form of the diffusivity equation.
3.11.3. Fetkovitch
Fetkovitch approximated the rate of influx into the reservoir using
an analogy of the productivity index
dWe
= J(p − pn+1 ), (3.11.4)
dt
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 200
where
J = the Fetkovitch influx constant,
p = the average aquifer pressure,
pn+1 = the pressure of the cell.
For finite aquifers, the influx into the reservoir is taken into account
to compute the average pressure by material balance for the next
time step.
2πkkro h
PI = (3.12.2)
µBo ln rrwe + S − 0.5
2πkkro h
PI = (3.12.3)
µBo ln rrwe + S
2πkkro h
P I∗ = (3.12.5)
µBo ln rrwe + S − 0.5
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 202
qo = P I ∗ (p − pwf ) (3.12.6)
where
qo = the volumetric flow rate of oil,
Twell = the well connection factor,
Mo (p, Sw ) = the oil mobility term,
pwf = the flowing bottom hole pressure,
ε = the gravity correction term.
The well connection factor (or well index) is defined as
αkh
Twell = , (3.12.8)
ln rrwo + S
where
α = the unit conversion factor,
kh = the permeability thickness,
ro = the pressure equivalent radius,
rw = the wellbore radius,
S = the skin factor.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 203
kro (Sw )
M (p, Sw ) = . (3.12.10)
µo (p)Bo (p)
For a high productivity index well with a low draw down, the
computed flow rate may be sensitive to the gravity correction, ε,
which corrects from the well reference depth to the sand face depth.
Various methods are available for this calculation depending on
the treatment of the multi-phase density and the dating of the
calculations. For very high deliverability wells, the results should be
checked for sensitivity to this correction.
3.12.2.1. Targets
Target rates are those rates which the simulator should attempt to
maintain provided that this can be achieved within the set operating
constraints.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 204
3.12.2.2. Constraints
Because a well in the simulation model could produce fluids in a
way impossible for a real well, the user can elect to choose some
constraints which can be applied.
These constraints may be related to wells, (for example a
minimum tubing head pressure for all wells corresponding to the
pressure needed to deliver hydrocarbons to the separator) or to
groups of wells (a group of wells producing into one separator has
a maximum group liquid rate determined by the separator size and
retention time).
The constraints can be connected in a tree structure so that at
each node (well, platform and field) constraints are honoured.
As well as gas–oil ratio limits and water cut limits, secondary
rate limits can be set which will permit the primary target rate to
be met provided that the secondary limit is not violated. This can
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 205
3.12.2.3. Actions
Typically these actions consist of attempting to control the produc-
tion of unwanted components such as water and gas by either altering
the well rate (reducing the rate of high water cut wells), performing
a work over on the well to squeeze off the worst offending perforation
(this has a 100% success rate in models) or the well can be shut in.
If a group of wells has spare production capacity, then at some
predefined drilling rate, wells can be added in specified locations in
order to increase the production rate to correspond to the available
facilities.
3.12.3.2. Prediction
During the prediction phase it is necessary that the production
constraints such as minimum tubing head pressure and maximum
liquid through put are implemented.
Figure 3.12.1 shows typical tubing lift curves for different water
oil ratios. If several layers are perforated each having different water
saturation, changing the bottom hole pressure can modify the water
cut and the convergent bottom hole (and therefore tubing head)
pressure must be found by an iterative procedure.
As in most iterative schemes, the closer the start value is to the
solution, the faster the convergence and so during a prediction it
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch03 page 207
is worth while setting target rates which are close to the expected
production rate.
It is a measure of the quality of the history match that the
model should proceed from the history match to the prediction phase
without experiencing instability or discontinuity.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 209
Chapter 4
History Matching
Deryck Bond
Senior Consultant for Kuwait Oil Company
4.1. Introduction
It may be appropriate to start by attempting to define “history
matching” in the context of reservoir modelling. One possible
definition, variants of which are commonly used, is as follows:
“History matching is the process of modifying the model input data
until a reasonable comparison is made with historical data.”
209
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 210
The scope for having multiple models in the above cases differs.
In the first example the appropriate style of model (lots of history and
lots of data) may be relatively deterministic. This, and the increased
efforts needed to history match a field with a lot of data, could make
having a single model a sensible choice. For the second example, the
emphasis on assessing uncertainty and the reduced effort that may
be associated with matching models with less data may make using
multiple models both more useful and more practical.
Structural model
Seismic data
Wellbore seismic
4D seismic
Core data RCAL/SCAL Geological /
Sedimentalogical studies petrophysical
Fault seal studies geocellular
Open hole log data model
Petrophysical studies/studies of contacts/FWLs
Geochemical data Up-scaled geo-
RFT data model
Pressure transient data
Cased hole logs (PLT/TDT)
ProducƟon data IniƟal simulaƟon
StaƟc surveys/SIBHP data model
Flowing surveys/FBHP data
Material balance/Aquifer characterisaƟon studies
Part reservoir simulaƟon studies
History matched
Pressure maps
simulaƟon model
Fluid movement maps
Studies of grid size/pseudo relaƟve permeability studies
CompleƟon integrity studies
Dynamic model
for predicƟons
The use of these data holds out the prospect of the initial geological
models being “closer to reality” than would otherwise be the case.
This work is potentially very time-consuming but there is the hope
that the resulting history-matched models will be more useful.
with values from, for example, RFT and Thermal Decal Time
(TDT) logs.
What effect does this choice of time intervals have on the
comparison of observed and simulated data? We may expect that
saturations would not change rapidly over short periods. Comparison
of observed and simulated saturations would not be influenced too
much by this choice of time intervals.
This may not be the case for pressures. Pressure communication
may be relatively rapid over inter-well distances. The time variation
(over periods of less than a month) can have a significant impact
on pressures at an off-set well. This may impact how RFT data
are modelled. We may want to represent the time variation of well
rates in more detail prior to the time when RFT data is available.
For similar reasons there would be a preference for comparing
RFT data with pressures at the exact time the data were acquired
rather than from simulator data stored at (for instance) monthly
intervals.
The data review would also give an estimate of the accuracy of the
observed data that is being compared to the simulation data. This
would involve:
Well Well
flowing shut-in
Low pressure
zone
High pressure
zone
The data required for a history match would place more emphasis on
historical data (e.g. old completions, history of mechanical problems)
than would be needed for continuing operation.
Non-
Initial pressure reservoir
gradient
Good
Sand
Depth
Poor
Sand
Oil
gradient
RFT Pressure
Initial pressure
gradient observed
Match 1
Depth
Match 2
Oil
gradient
RFT Pressure
involving matching RFT data. Figure 4.5.2 shows some RFT data
taken from the oil leg of a reservoir. Next to the RFT data there is
an indication of reservoir quality.
There are a number of considerations we need to address when
deciding what a good match to the data could be. These include:
The aims we have in matching the data could influence what would
be considered a better fit to the data. Consider the two potential
“matches” to the RFT data shown in Fig. 4.5.3.
If we were interested in getting a good match to the average
pressure decline, then Match 1 would be preferred. If we wanted a
match that honoured the pressure break, then Match 2 might be
preferable.
where the wi are weights and the σi represents a target level for
fitting the data. Clearly in using such an approach care needs to be
taken in choosing appropriate wi and σi values.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 227
This only measures how well or badly observed data are matched.
We may also want to account for how realistic or unrealistic the
changed input parameters to the simulation model are. This is
discussed by Schulze-Riegert and Ghedan (2007).
If a perfect match is achieved then all of these methods will give the
same result because production of all the phases (and pressure) will
be matched. The choice of control mode should be dictated by the
way this influences our ability to match the model.
There are two arguments in favour of using reservoir volume
rates. The first relates to matching the reservoir pressure. Consider
the material balance equation for a reservoir (or for a fault block
within a reservoir) and for convenience assume there is no free gas.
The material balance equation is:
0.9
0.8
0.7
0.6
Water Cut
0.5 TRUE
Qres
0.4 Qoil
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5
Time
• simulated pressures are a little too high above the shale barrier;
• there is a decline of pressure below the shale barrier that cannot
be easily explained.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 235
Water
cut
Time
cut could indicate higher flow in the first layer to produce water.
Secondly the simulation model produces water earlier than is actually
the case.
When we seek to match the water cut we need to think about
the other data that are available, e.g.
• PLT data that could give information of flow in different layers;
• PNL data that could give information of swept zone saturations.
What changes could be made to better match the water cut
development?
The split kh product and ϕh product will influence the style of
water cut development, as will the mobility ratio of the displacement.
The timing of water production will be influenced by volumetric and
local sweep efficiencies.
One approach to modelling that data could be as follows.
• If PLT data are available then attempt to match the split
between layers by changing modifying layer permeability values
(but keeping total kh in line with estimates from PBU data).
• Investigate how reasonable changes to reliable permeability end
points and the shape of reliable permeability curves could influence
the time of water production. The following could give later water
production:
◦ decreasing Sorw (and hence decreasing local sweep efficiency);
◦ decreasing Krw — this would make the displacement less stable
(it would tend to reduce volumetric sweep and local sweep
efficiencies).
• Investigate how reasonable changes to permeability and porosity
could change water production. The following could give later
water production:
◦ reducing permeability heterogeneity within the main flow units;
◦ possibly increasing kv/kh;
◦ increasing porosity.
Based on this work we can try to produce a reasonable match to the
data. There are two potential problems.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 240
Sealing
fault?
of matching flowing BHP and water cut then this matching should
involve making at most rather minor changes to tubing performance
curves.
4.7. Conclusions
Producing history-matched models may allow us to address impor-
tant questions on reservoir development and management. The
quality and utility of the models depends on how well dynamic
data are accounted for in the modelling process. Much of the effort
associated with accounting for the dynamic data should form a
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-ch04 page 244
References
Mattax, C. C. and Dalton, R. L. (eds.) (1990). Reservoir Simulation, SPE
Monograph Series, Volume 13, Richardson, Texas.
Peaceman, D. W. (1978). Interpretation of well block pressure in numerical
reservoir simulation, SPE Journal, 18(3), 183–194.
Schulze-Riegert, R. and Ghedan, S. (2007). Modern Techniques for History
Matching, Proceedings of the 9 th International Forum on Reservoir
Simulation, 9–13 December, Abu Dhabi, UAE.
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-index page 245
Index
245
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-index page 246
Index 247
K mixed wet, 57
Kozeny–Carman equation, 14 mobility, 37
mobility control, 66
L mobility ratio, 61, 93
model capillary pressure, 174
layered, 15, 17
model components, 134
layered reservoir, 45
model control, 134
layered rock, 14, 16
model fluid properties, 190
Leverett, 30, 44
model relative permeability, 161
light tight oil, 52
model well and production data, 201
limestones, 12
Mukerji, 44
linear solvers, 126–127, 130
multi-contact miscibility, 75
liquid, 19
multi-contact miscible gas, 103
local grid refinement, 144–145, 147
multi-contact miscible processes, 74
log data, 148, 157
multilateral, 66
logging, 36
multiple models, 211
Lorenz coefficient, 66
low salinity, 49, 103
N
low salinity water, 50, 101
low salinity water flooding, 92, 97 non-orthogonality, 147
non-wetting fluids, 56
M non-wetting liquid, 24
non-wetting phase, 33
macroscopic sweep, 63, 65–66, 103
nonlinearity and outer iterations,
macroscopic sweep efficiency, 59
126
Magnus, 102
numerical aquifer, 200
manipulation of capillary pressure,
numerical models, 112
176
numerical resolution, 217
mass conservation, 113–114, 117
matching fluid movement, 236
O
matching pressure, 233
matching water movement, 238 Ohm’s law, 36–38
material balance calculations, 42–43 oil, 1–4, 6–7, 21–22, 25–26, 28, 30,
mathematical models, 110, 112 32–36, 40–41, 45–46
Mavko, 44 oil supply, 53
method of characteristics, 81, 84 oil wetting, 57
microscopic displacement, 67 oil–water contact, 26, 28
microscopic displacement efficiency, oil–water transition zone, 27
59, 94, 97, 103 oil-wet, 25
microscopic displacement factor, 58 oil-wet relative permeabilities, 99
microscopic fluid distribution, 56 oil-wet rocks, 40
mineral grain volume, 42 optimum WAG ratio, 90
mineral volume, 2 output, 115, 135
minimum miscibility pressure, 101
miscible, 103 P
miscible displacements, 60 parallel-tube model, 27, 34, 39, 45–46
miscible gas, 101 part field models, 221
August 11, 2017 10:20 Topics in Reservoir Management - 9in x 6in b2764-index page 249
Index 249
Index 251