Вы находитесь на странице: 1из 19

September 2010 SPE Journal 751

Formation-Damage
and Well-Productivity Simulation
Arild Lohne, SPE, and Liqun Han,* SPE, International Research Institute of Stavanger;
Claas van der Zwaag, SPE, Statoil; Hans van Velzen, Nederlandse Aardolie Maatschappij (NAM);
Anne-Mette Mathisen, SPE, Statoil; Allan Twynam, SPE, BP; Wim Hendriks, SPE, Shell; Roman Bulgachev, SPE, BP;
and Dimitrios G. Hatzignatiou, SPE, International Research Institute of Stavanger
*Formerly at the International Research Institute of Stavanger (IRIS), now with Halliburton.
Copyright 2010 Society of Petroleum Engineers
This paper (SPE 122241) was accepted for presentation at the 8th European Formation
Damage Conference, Scheveningen, The Netherlands, 2729 May 2009, and revised
for publication. Original manuscript received for review 3 April 2009. Revised manuscript
received for review 11 December 2009. Paper peer approved 23 December 2009.
Summary
In this paper, we describe a simulation model for computing the
damage imposed on the formation during overbalanced drilling.
The main parts modeled are filter-cake buildup under both static
and dynamic conditions; fluid loss to the formation; transport of
solids and polymers inside the formation, including effects of pore-
lining retention and pore-throat plugging; and salinity effects on
fines stability and clay swelling. The developed model can handle
multicomponent water-based-mud systems at both the core scale
(linear model) and the field scale (2D radial model). Among the
computed results are fluid loss vs. time, internal damage distribu-
tion, and productivity calculations for both the entire well and
individual sections.
The simulation model works, in part, independently of fluid-
loss experiments (e.g., the model does not use fluid-leakoff coef-
ficients but instead computes the filter-cake buildup and its flow
resistance from properties ascribed to the individual components
in the mud). Some of these properties can be measured directly,
such as particle-size distribution of solids, effect of polymers on
fluid viscosity, and formation permeability and porosity. Other
properties, which must be determined by tuning the results of the
numerical model against fluid-loss experiments, are still assumed
to be rather case independent, and, once determined, they can be
used in simulations at altered conditions as well as with different
mud formulations. A detailed description of the filter-cake model
is given in this paper.
We present simulations of several static and dynamic fluid-
loss experiments. The particle-transport model is used to simulate
a dilute particle-injection experiment taken from the literature.
Finally, we demonstrate the models applicability at the field scale
and present computational results from an actual well drilled in the
North Sea. These results are analyzed, and it is concluded that the
potential effects of the mechanistic modeling approach used are
(a) increased understanding of damage mechanisms, (b) improved
design of experiments used in the selection process, and (c) bet-
ter predictions at the well scale. This allows for a more-efficient
and more-realistic prescreening of drilling fluids than traditional
core-plug testing.
Introduction
A simulation tool, referred to as Maximize, has been developed
for the purpose of investigating fluid loss to the formation during
overbalanced drilling and the impairment imposed on the forma-
tion by the invading fluid. The objective of the program is to serve
as a tool for supporting well planners decisions related to the
choice of well fluids, and integrating, analyzing, and interpreting
laboratory and field formation-damage data.
The filtration properties of the mudcake, forming at the wellbore
surface, have been investigated by several authors over the years;
see, for example, Ferguson and Klotz (1954), Outmans (1963),
Bezemer and Havenaar (1966), Arthur and Peden (1988), Fordham
et al. (1991), and Dewan and Chenevert (2001). The common
understanding is that, once the filter cake has been formed, it will
control the filtration rate independently of the formation properties,
except at very low permeability where the flow resistance offered
by the formation is comparable to the filter-cake resistance.
The properties of the filter cake depend only on its composition,
the pressure drop over the cake p, and the shear stress acting on the
cake surface by the circulating mud. Under dynamic conditions, the
filtration rate will approach asymptotically a limiting steady-state
rate, which depends only on the shear stress at the cakes surface.
Semmelbeck et al. (1995) combined a filter-cake-building-
and-filtration model with a fluid-flow simulator for computing
the fluid-invasion profile along the well. Further improvement
of the filter-cake model was presented by Dewan and Chenevert
(2001), who demonstrated the derived models capability to repro-
duce complex laboratory experiments with sequential changes in
dynamic shear rate and overbalance pressure. Others also have
investigated filtrate invasion by numerical simulations [e.g., Ding
et al. (2002), Wu et al. (2004), and Suryanarayana et al. (2007)].
Knowledge of the filtrate profile along the wellbore is important
for well log interpretations and for evaluating the damage imposed
on the formation by the invading fluid. A number of potentially
damaging mechanisms and modeling of these are described in
detail by Civan (2000). For water-based muds, one of the main
damage mechanisms is particulate plugging of the formation,
either by externally introduced particles or by in-situ fines mobi-
lized by the invading fluid. The mechanisms of particle transport in
porous medium have been investigated by Gruesbeck and Collins
(1982), Sharma and Yortsos (1986), Wennberg and Sharma (1997),
and Al-Abduwani et al. (2005).
Traditionally, the computation of fluid loss has relied on using
experimental filtration rates or, as in the more-sophisticated mod-
els, experimental cake permeability and porosity obtained for a
specific mud. In our approach, we bring this one step forward by
computing the cake properties from its composition. The proper-
ties from which the cakes porosity and permeability are computed
are ascribed to the individual components. In Ding et al. (2002)
and Suryanarayana et al. (2007), the damage in invaded zones is
simulated on the basis of experimental return permeability and
an assumed damage profile. We make no such assumptions but
model the flow and retention of both particles and polymers inside
the formation and compute the permeability reduction from the
amount of trapped material.
Models that are used should be consistent with the physics
involved at a macroscopic level. Input parameters should ideally
be of three kinds: (a) properties of involved components (e.g.,
viscosity, density, size of particles), (b) properties of the formation
(e.g., permeability, porosity), and (c) model parameters describing
events that are independent of Parameters a and b. The benefit of
this approach is that the amount of empirical input parameters is
reduced and the ones used are more universal and show less vari-
ance among different experiments. Finally, the proposed approach
will increase the use of earlier experience because, ideally, all
experiments should be matched by a single data set, provided
adequate descriptions of the experiments are available.
752 September 2010 SPE Journal
Simulation Model
This section describes the various modules that constitute the newly
developed simulation model. The two main parts of the Maximize
program are (1) the filter-cake model handling filter-cake buildup
and controlling flow into the formation and (2) the reservoir-flow
model handling flow inside the formation. The fluids introduced to
the formation contain a number of dissolved and dispersed com-
ponents, which, in turn, may change the original flow properties
of the formation through various chemical and physical processes.
The retention of solids and polymer is split into a pore-throat-trap-
ping model and a pore-lining-adsorption model. Brine interaction
with the rock surface and clays is described by a multicomponent
cation-exchange model, which provides input to processes such as
fines migration and clay swelling.
Flow Model. The ow model handles two-phase ow in 1D
linear (typically used for cores), 2D rectangular, and 2D radial
models (appropriate for the eld scale). We use two boundaries
(wells) constrained by constant rate or constant pressure. The ow
equation is solved using the implicit-pressure/explicit-saturation
(IMPES) method (implicit in pressure and explicit in saturations
with respect to time); see Aziz and Settari (1979) for further details
on reservoir simulators.
The phase flow between grid cells is computed from the
pressure solution. Knowing the flow, local concentrations of oil,
water, and dissolved or suspended species are updated for the new
timestep. Chemical and physical processes are decoupled from the
flow equation. Components are first moved between cells, and
then changes over the timestep caused by physical and chemical
processes are computed locally in each grid cell.
Rock properties, such as absolute permeability k, porosity
, relative permeabilities, and clay content are allowed to vary
between layers (sections). Moderate compressibility of both the
formation and the fluid components is handled by the model.
Corey-type relative permeabilities are used.
k k S S
S S
S S
j o w
rj rje jn
Ej
jn
j jr
wr or



, , ,
1
. . . . . . . . . . . . . . . . . (1)
As stated earlier, this is a multicomponent simulator. Properties
of components, such as viscosity or component effect on viscosity,
are described in the following sections.
In simulations with mud filtration, the rate will change very
rapidly during the first period when the filter cake is built up from
zero thickness, and an IMPES-type solution will be unstable. We
solve this by precomputing the growth of the cake during the
timestep, assuming the pressure difference between the well (mud
pressure) and the connection block to be constant. The estimated
cake properties are used in the computation of the new pressure
solution for the total grid, and, finally, the rates from this solution
are used to update the real growth of the filter cake.
Filter Cake. The lter-cake model presented here computes cake
properties from properties of the individual components present
in the mud. The model handles an arbitrary number of polymers
and solids.
Static Filtration. The relation between cake thickness h
c
and
filtrate volume V
f
is obtained from mass balance (Bourgoyne et al.
1986):
h
V C
A
C
c
c c
c
f
c
sm
sc sm

1
1
, , . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
where A
c
is the filter-cake area and c
sm
and c
sc
are the concentrations
of solids in the mud and in the filter cake, respectively. The flow
rate through the cake expressed by Darcys law is
q
pA k
h
c c
c

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
Setting the rate q dV
f
/dt, inserting h
c
from Eq. 2, and integrat-
ing the resulting expression yields the known-filtrate-volume vs.
square-root-of-time relation (a spurt-loss volume V
sp
is also added
to the resulting expression):
V V A C t C
pk
C
f sp c st st
c
+ ,
2
1

. . . . . . . . . . . . . . . . . . . . (4)
There are three unknowns in Eq. 4: the spurt loss V
sp
, the cake
permeability k
c
, and the cake porosity
c
(c
sc
1
c
is part of C
1
).
Two of these variables, k
c
and
c
, depend on the mud composition
and the pressure drop over the cake, while the spurt loss depends
on the mud composition and properties of the rock. The cakes
porosity and permeability will decrease with increasing pressure
drop because of the compressibility of the cake and thus will reduce
the rate increase.
The main effect of temperature is through its influence on the
filtrate viscosity and thereby on the filtration parameter C
st
. Dif-
ferent polymers used in mud will have different upper temperature
limits above which they become chemically unstable. Below this
limit, only a weak temperature dependency is assumed.
Cake Permeability. The permeability of a porous medium can be
estimated with the Carman-Kozeny equation (Lake 1989). Because
mud is a multicomponent system containing different kinds of solids
of very different sizes, we choose the following variant:
k
S
S
S c
c
i i
i

( )


3
2
0
2 0
0
2 1
, . . . . . . . . . . . . . . . . . . (5)
The tortuosity parameter (L
t
/L)
2
accounts for the effective
flow-path length L
t
, c
i
is volumetric concentration of species i,
and S
0
is the specific surface area of the medium defined as the
surface-to-volume ratio. In particular for spheres, S
0
6/D
p
, where
D
p
is the particle diameter. The specific surface area in a mixed
system is readily obtained by volume-weighted averaging. The
operator <> used for S
0
indicates that the property represents a
volume-weighted average.
Another useful expression that relates an effective pore diam-
eter to permeability and porosity is
D
k

32
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
The Carman-Kozeny approach treats the porous medium as a
bundle of capillary tubes, and D

corresponds to the tube diameter


when all tubes have the same diameter.
Cake Compressibility. The compression of the cake is a result
of the local pressure difference between the cake matrix and the
fluid filling the pore space. The cake-matrix pressure is assumed
constant and equal to the mud pressure p
m
. The fluid pressure
decreases through the cake as a result of the viscous flow. Because
this pressure difference will vary throughout the cake, physical
cake properties such as porosity and permeability, which depend
on the overburden pressure, will be functions of spatial position.
We are interested in some effective average properties, which we
will obtain by integrating over the cake thickness.
The local pressure gradient within the filter cake can be described
by Darcys equation. At a given time, the flow rate u will be constant
through all layers of the cake for an incompressible fluid. Then,
according to Darcys equation, the term k
c
dp/dx must be constant
(independent of location in the cake) and equal to u. (assuming
is constant).
d
d
d
d
p
x
p
x
u
k x
n
c


( )
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
Let p
n
represent the local net overburden pressure in the cake
(p
n
p
m
p), and let p represent the fluid pressure drop over the
cake. The overburden pressure p
n
will range from zero at the cake
inlet to p at the cake outlet. The velocity u can be expressed in
September 2010 SPE Journal 753
terms of the average cake permeability k
ca
and p. Let k
c
be a
function of p
n
, and combining k
ca
with Eq. 7 gives
k p p u x u
k p
h
k
k p
c n n
h p
ca
c
ca
c n
c
( )

(

d d

0 0

,
))

dp
p
n
p
0

. . . . . . . . . . . . . . . . . . . . . (8)
The effective permeability will be a function only of p and
will be independent of the cake thickness. To solve the integral
of k
c
(p
n
), we need a functional relation between k
c
and p
n
. In the
literature, k
c
and
c
are normally treated uncoupled using the same
equation type (Outmans 1963):
k k p p n n
c o
n
c o
n


1 2
1 2
0 and , , . . . . . . . . . . . . . . . . . . (9)
In our case, we want to calculate the permeability using the
Carman-Kozeny approach of Eq. 5, include the effect of compress-
ibility through the porosity, and integrate k
c
dp
n
to obtain the aver-
age k
ca
. We use the classic definition for compressibility:
n
p
e n
c o
n p
c
c n

1
0


d
d
, . . . . . . . . . . . . . . . . . (10)
Note that the porosity in Eq. 10 converges to the correct limit

o
when p
n
approaches zero, which is not the case with Eq. 9. The
solution to the average cake permeability, as well as the computa-
tion of the average porosity, is given in Appendix A. Here, we show
only the final expressions. Defining z(p
n
) 1(p
n
), the average
permeability of a compressible filter cake is
k
S n
F p F
p
F p
z
z
ca
c
k k
k n

( ) ( )
( ) +
1
2
0
1
2
0
2

,
ln zz . . . . . . . . . . . . . . . . . . . . . . . (11)
and the average filter-cake porosity is

( ) ( )
( ) ( )
( ) +
F p F
F p F
F p
z
z
k k
n

0
0
1
3 3
,
ln zz z +
1
2
2
. . . . . . . . . . . . . . . . . . . . . . . . (12)
The average filter-cake permeability and porosity computed by
Eqs. 11 and 12, respectively, can be used directly in Eqs. 2 and 4.
The compressibility of the filter cake can be split into a revers-
ible (elastic) and an irreversible (compaction) part. A simple way to
model this is to apply an irreversibility factor, F
irf
. Let p
max
be the
historical maximum p. If the current pressure drop p is less than
p
max
, then an apparent effective pressure drop p
a
is used:
p
a
p + F
irf
(p
max
p). . . . . . . . . . . . . . . . . . . . . . . . . (13)
Partially reversible compression of a filter cake is illustrated
in Fig. 1.
Solids. The information needed in the filter-cake model includes a
size distribution of the solid particles, the specific surface area S
0
, and
the solids contribution to the cakes porosity and compressibility.
The size distribution can be either in the form of a log-normal
distribution or entered as a table. If a tabular form is used, the
volumetric cumulative distribution function F [0, 1] is taken to
be piecewise linear when plotted against ln(D). S
0
is computed by
integrating S
0
(D) over the size distribution. If tabular input with
n rows is used,
S D
D
S
e
c x c
D
sh
i
sh
x
x
x
i sh i
i
i
i
0
6
6
6
1
1
( )

,
d

j
(
,
\
,
(

( )
1
1
1
0
1
D
c
F F
x x
x D
S
i
i
i i
i i
,
, ln ,


S
i
n
2
. . . . . . . . . . . . . . . (14)
The shape factor
sh
is defined as the ratio of the particles surface
area to the surface area of a sphere with the same volume (i.e., for
spheres,
sh
1). Examples for other shapes:
sh
is 1.24 for a cube
and 2.3 for a flat particle with dimensions 0.111 in any units.
For each solid type present in the mud, a reference porosity
o

corresponding to zero overburden pressure and a compressibility
parameter n
c
are required.
Polymer. For polymers, we need to know their effect on fluid
viscosity, which is important in dynamic filtration and for flow
inside the porous medium. Second, we need the size of the poly-
mer molecules to evaluate entrapment in the filter cake and in the
formation. Finally, we must estimate the resistance to flow caused
by the polymer in the filter cake.
Polymer Viscosity. The effect of polymer concentration on the
fluid viscosity is typically expressed at low shear rate with a modi-
fied Huggins equation (Sorbie 1991). We use the expression

0
2 2 3 3
1 + + + ( )
s
c k c k c , . . . . . . . . . . . . . . . . . . . . (15)
where c is the polymer concentration (volume fraction),
s
is the
solvent (e.g., water) viscosity, and is a dimensionless form of the
intrinsic viscosity [] in units [cm
3
/cm
3
]. The intrinsic viscosity [] is
defined as the limit of the reduced viscosity when polymer concentra-
tion approaches zero. k is the Huggins constant, which, for a range
of polymers in good solvents, is reported to be equal to 0.4 0.1.
Eventual changes in temperature and in molecular weight M
w
caused
by, for example, selective trapping of larger polymer molecules in
pore throats are accounted for by the following expression:
[ ]
( ) ,

,
]
]
]
( ) ,

]
] p
w
w
a
vT
M x t
M
B T T
,
0
0
1 , . . . . . . . . . . . . . . (16)
where exponent a is reported to be in the range 0.51 (higher value
for good viscosifiers). T is the actual temperature, T
0
is a reference
temperature at which [] is measured, and B
T
is an empirical
parameter. The units for the density
p
must be consistent with the
units used for [] so that becomes dimensionless. The shear-thin-
ning effect of polymers is represented by Meters model (Meter and
Bird 1964), with the final polymeric viscosity given by
0.3
0.35
0.4
0.45
0.5
0.55
0.6
0.65
0.7
0 5 10 15 20 25
Cake Overburden Pressure, bar
P
o
r
o
s
i
t
y
1st curve
F
irf
=0.2
F
irf
=0.8
n
c
=0.04
Fig. 1Compressible filter-cake porosity, calculated with
o
=
0.65, n
c
= 0.04 bar
1
, and irreversibility factor F
irf
= 0.2 and 0.8.
754 September 2010 SPE Journal

p s
s
hf
P
+

+
j
(
,
\
,
(

0
1
1

, . . . . . . . . . . . . . . . . . . . . . . . . . . . . (17)
where exponent P has a value between 1 and 2 (high shear
thinning) and
hf
is the shear rate at which (
0

s
) is reduced
to one-half. The main influence of temperature is through the
solvent viscosity
s
. We use the same Arrhenius-type equation for
oil and water to compute the viscosity as a function of pressure
and temperature:


j
(
,
\
,
(
+ ( )
,

,
,
]
]
]
]
o
B
T T
B p p
e
T
o
p o
1 1
, . . . . . . . . . . . . . . . . . . . . . . . . . . . (18)
where
o
is the reference viscosity at temperature T
o
and pressure p
o
.
B
T
and B
p
are empirical constants, and temperature is in absolute units
(K). We use the fluid viscosity in Eq. 18 to describe flow through the
filter cake (Eqs. 3 and 4). Yet another correction for the content of
solids is made before the effect of polymer is computed:


s
o
sm
c

( ) 1
2 5 .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (19)
Polymer Size. We assume a log-normal distribution of M
w
in the
computations. The size of the polymer molecule in solution can be
estimated from the intrinsic viscosity [] using Florys empirical
equation for the mean end-to-end distance:
D M
hp w
[ ] ( ) 0 00017
1 3
. . . . . . . . . . . . . . . . . . . . . . . . . . . (20)
The molecular weight M
w
is in g/mol, [] in mL/g, and the com-
puted polymer diameter in m. We will use the hydraulic polymer
diameter D
hp
for computing polymer entrapment in the filter cake
and pore-throat trapping of polymer inside the formation.
Flow Resistance. The use of the Carman-Kozeny approach of
Eq. 5 requires the polymers contribution to the specific surface
area and the cake porosity. The surface of the polymer is estimated
by assuming the molecule to be a long rod with diameter D
ch
. The
specific surface area is then
S
D
p
ch
0
4
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (21)
Note that the specific surface area for a given polymer type
depends on the chain diameter D
ch
of the repetitive unit but not
on the molecular weight.
For each polymer type i, we specify a
oi
and n
ci
according to
Eq. 10 that will go into the computations of the effective param-
eters for the cake.
Cake Buildup. The cake-buildup calculations involve two peri-
ods, the spurt-loss period and the filtration period after a filter cake
has been formed.
Spurt-Loss Period. In the spurt-loss period, all polymers and
solid particles smaller than a critical size D
crs
are assumed to enter
the formation. Solid particles larger than D
crs
will be held back on
the rock surface. When sufficient coverage of the rock surface is
reached, all solids will be retained. The critical particle size for
solids is computed from
D D
crs s


, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (22)
where D

is obtained by Eq. 6 and parameter


s
represents the
lower size ratio of particle size to pore size for particles that will
be retained at the rock surface. A common rule of thumb is that
particles larger than 1/3 of the average pore size will form an
external filter cake (Pautz et al. 1989; van Oort et al. 1993). This
rule indicates a critical size where there is a high probability for
blockage at the rock surface over an extended time period, during
which several particles will pass before the blockage occurs. The
meaning of
s
is slightly different; it represents the average low-end
size of particles held back on the rock surface during a short spurt-
loss period. Note also that D

is an average property representing


both pore-body and pore-throat distributions. Some particles larger
than D
crs
will enter the rock, and some smaller particles will be held
back. Knowing the particle distribution and assuming an initial cake
thickness (see next paragraph), the value of
s
can be derived from
experimental spurt-loss values. Our experience so far indicates that
a value for
s
in the range of 0.30.5 seems to be appropriate. The
effect of previous internal permeability reduction is included in the
computed pore diameter D

. This means also that particle distribu-


tions smaller than initial D
crs
will eventually form a cake when the
internal permeability is sufficiently reduced.
The average height of the layer needed to form a covering layer
on the rock surface is computed by
h n h h
S D D
sp sp crs crs
crs

> ( )
,
6
0
, . . . . . . . . . . . . . . . . . . . . . (23)
where h
crs
represents a characteristic height of the retained particles
computed from the specific surface area for the fraction having D >
D
crs
. n
sp
is the number of layers needed to form a completely
covering cake. Note that the proposed way of computing h
sp
place
more weight on the smaller particles and reduces the spurt loss for
flattened particles. The spurt loss volume is then
V A
h
C
C
c
c c
sp c
sp
sp
sp
sm
sc sm

1
1
,
*
* *
, . . . . . . . . . . . . . . . . . . . . . . (24)
where c
sm*
is the concentration of solids with D > D
crs
in the mud
and c
sc*
is equal to (1
o
) computed for solids alone. In the special
case of spherical and monosized particles, n
sp
1 will correspond
exactly to a single monolayer with porosity
o
. For n
sp
, a value in
the range 12 is suggested.
Filtration Period. In this period, all solids are retained in the
cake. Material deposited during the spurt-loss period is ignored
(i.e., cake thickness starts from zero). The first mechanism to
evaluate is retention of polymer in the cake. The part of the poly-
mer molecules with D
hp
smaller than a critical size D
crp
will pass
through the cake. D
crp
is estimated as a fraction
p
of the pore size
D
c
(Eq. 6) for a filter cake made up of solids alone:
D D
crp p c


. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25)
The effective polymer concentration that will be retained in
the cake is then obtained from the M
w
size distribution. The criti-
cal size ratio
p
represents the average probability for entrapment
over a long filtration period and should be in the low range of
standard trapping rules (1/71/3). A value of 0.1 has been used in
the simulations presented here.
Finally, the cake properties can be computed. First, effective
cake porosity, compressibility, and specific surface area are esti-
mated as the volume-weighted average of all involved components,
solids, and polymers (n
t
components):

o
c
c i o i
i
n
c
c
c i c i
i
n
c
c
n c n
S c
t
t

, ,
, ,
,
,
1
1
0 cc i i
i
n
c i
i
n
S
c
t
t
, ,
,
,
0
1
1
1

, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (26)
where c
c,i
represents the relative concentration of Component i in
the filter-cake matrix. The average cake permeability and porosity
are obtained readily from Eqs. 11 and 12, respectively, and, finally,
the static filtration parameter C
st
in Eq. 4 can be computed.
September 2010 SPE Journal 755
Dynamic Filtration. The filter-cake model outlined so far applies
for static conditions at the filter-cake surface. In real situations, the
filter-cake surface is exposed to shear forces from mud circulat-
ing through the well. Two main mechanisms are considered in the
dynamic filtration model, namely, reduced attachment and erosion.
The flow along the cake surface will cause part of the particles
to roll off the surface and bounce back into the mud. Only a fraction
of the particles will stick to the cake surface. This will reduce the
growth rate of the cake. The shear stress at the cake surface may
also detach particles from the filter cake and, thus, decrease the
cake thickness. This is called the erosion rate Y. The cake growth
in terms of and Y can be obtained from mass balance:
d
d
h
t
C u
Y
c
C
c
c c
c
d x
sm
d
sm
sc sm

j
(
,
\
,
(

1
1

,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . (27)
The flow through the cake u
x
is given by Darcys law if the
cake permeability and current thickness are known. This leaves
two unknowns to be determined, and Y.
Consider a nonmoving spherical particle just in contact with
the filter cake surface with a filter loss rate u
x
and a perpendicular
mud flow along the surface u
z
. For simplicity, we assume a constant
shear rate near the surface, so that the average velocity u
z
in a
layer with the same thickness as the particle will be
d
d
,
d
u
x
u x x
u
D
x x D
z
z
z
p
D
p
p

( )


1
0
1
2
. . . . . . . . . . . . . . . . . . . . . . . . . . . . (28)
The force acting on a spherical particle with constant relative
velocity in a fluid is given by
F uD
p
3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (29)
Then, for the particle close to the surface, the forces in the x
and z directions, respectively, are
F u D
F D
x o x p
z p

( )
3
3
2
2

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (30)
Note that, in F
z
, the viscosity will be a function of the
shear rate while the viscosity acting toward the surface (in F
x
) is
computed at zero shear rate (
o
). The dominating mud flow will
be along the surface. The reduced viscosity at higher rates (shear
thinning) is caused by polymer molecules orienting themselves
along the main flow direction. Particles bouncing back into the
mud will not see this shear thinning because they will be moving
perpendicular to the flow. The ratio of forces acting on the particle
along and toward the surface is then
F
F
F
D
u
R
z
x
p
o x

( )


2
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (31)
The fraction of solids (in the mud) that will attach to the surface
() is modeled by

+
( )
1
1 F f
R n
n
F
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (32)
The exponent n
F
determines how fast the transition of a is when
the filtration rate u
x
decreases, and its value should reflect the width
of the particle distribution. Values in the range 0.51 are used in
the simulations here. Finally, f
n
is a friction factor.
The shear stress acting on the filter cake surface is given as


z z
z

( )

,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (33)
Assume a multilayered cake, with layer thickness D
p
(a char-
acteristic particle diameter size). The forces holding the layer in
place are (1) cohesive attraction between the particles (solids and
polymers) that form the cake matrix and (2) the viscous pressure
gradient from the flow through the cake. We assume cohesive
forces to be represented by the filter cake yield strength
y
, which
is measured by Cerasi et al. (2001) to range from a few hundreds
to a few thousands of Pascals. The pressure drop over the first layer
will be equal to the net overburden pressure at distance D
p
into the
cake, p
n
(x D
p
), which we obtain from Eq. A-6. The x-directed
forces considered are then

x n p y
p x D
( )
+ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (34)
We assume the erosion rate to be proportional to the ratio of forces
acting in the z and x directions and to D
p
. An empirical rate constant

e
(in units s
1
) is introduced. The erosion rate is given by
Y D
p D
e p
z
n p y

( )
+

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . (35)
The particle diameter used in Eqs. 28 through 35 is estimated
from the size distribution of both solids and polymer using vol-
ume-weighted D
1
.
A limitation of the filter-cake model is that we are using average
cake properties. This prevents us from properly including physical
processes such as size-selective attachment to the cake and pressure-
dependent yield strength
y
. A reasonable implementation of such
effects requires some kind of spatial discretization of the filter cake.
Particle Retention in the Formation. Retention of particles ow-
ing through a porous medium can be divided into two main parts,
pore-throat plugging and pore-lining retention. The retention kinet-
ics for the two parts is very different and so is the damaging effect.
In the case of external particles (injection of suspended particles),
pore-throat trapping will be most important close to the inlet, while
pore-lining retention will dominate deeper in the core. Larger par-
ticles will be ltered out from the solution as they are trapped in
narrow pore throats, while smaller particles can penetrate deeper.
In laboratory experiments, normally only the combined effects of
the two processes are obtained, resulting in a large variety of reten-
tion models/parameters. The observed results will strongly depend
on the experimental setup, and there is no good method for scaling
the experimental results to other conditions.
Here, we will try to separate the two retention types. Our main
focus is on the pore-throat trapping because it has the higher dam-
age potential.
Pore-Throat Trapping. In pore-throat trapping, only particles
large enough to be held back in the narrow parts of the pore space
(throats) will be involved. To model this, we need to keep track of
particle sizes. Solid and polymer components are split into a number
of size fractions, each represented as a monosized subcomponent in
the simulation. The basic equation used for both solids and polymers
(k s, p) allows for both trapping and re-entrance of particles back
to solution. The trapping rate for Subcomponent i is computed by
d
d


i w
w
k t i k i
t
u
S
c
( )
1 2 , , ,
, k s, p. . . . . . . . . . . . . . . . . . (36)
The units for trapped material and total mobile concentration of
Component i are pore volume (PV) units, so that, over a timestep t,
we have
i
c
t,i
. The unit for parameters
1
and
2
is 1/length.
For solids, the detrapping term
2
will be (approximately) zero. If
2

is zero, Eq. 36 says that a fraction of all passing particles equal to

1
will be retained per length unit and the expected traveling length
756 September 2010 SPE Journal
E[x] is 1/
1
. To make the equation more general, we allow
1
to be
a function of particle size and
2
to depend on fractional flow. For
solids, if the subcomponent size D
i
>
ts
D

,
1
and
2
are given by

1 10
2 20 21
11
1
s s
i
s s s o
D
D
f
s

j
(
,
\
,
(
+ ( )
,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (37)
If the particle size is smaller than this limit, trapping is not
considered. Finally, D

is the pore size estimated by Eq. 6.


Contrary to rigid solid particles, polymer particles will be quite
deformable, and their retention in pore throats may be highly
dependent on the viscous drag from fluid passing through. For
polymers,
1
and
2
are computed by

1 10
2 20 21
11
1
p p
hp i
p p p o
d
D
f
p

j
(
,
\
,
(
+
(
,
,
))
+

j
(
,
\
,
(
,

,
,
]
]
]
]
1
22
22
p
p
p

. . . . . . . . . . . . . . . . . . (38)
The size of the polymer Subcomponent i is represented by the
hydraulic diameter from Eq. 20, and a lower size limit for pore-
throat trapping can be specified by
tp
D

. Particles that are smaller


than this limit will not be trapped in the simulations. The main dif-
ference from solids is that the re-entrance term will increase with
increasing pressure gradient p and cannot be ignored.
Two methods for computing the permeability reduction caused
by the pore-throat trapping are implemented. In the first, an inter-
nal-cake permeability k
ic
is computed similarly to k
c
for the filter
cake, except that a zero overburden pressure p
n
is used. Then, the
effective absolute permeability is computed by harmonic averaging
between k
ic
and the original rock permeability k:
k
a
k
a
k
a
ic
s p
ic
eff
+
j
(
,
\
,
(

+

1
1
1
,

. . . . . . . . . . . . . . . . . . . . (39)
The internal-cake porosity is a weighted average <
o
> of all
trapped material. This method seemed to overpredict the perme-
ability reduction, in particular when polymer is involved. The
likely explanation is that small amounts of trapped material will
not be packed close together as a cake, but rather will be spread
over a larger area within the narrow parts of the pore space. One
way to compensate for this is to introduce a modification factor
for the specific surface area of the polymer f
pm
.
In the second method, we compute a pore-throat permeability
k
pt
within the narrow pore space defined as a fraction f
pt
of the total
pore space. A specific surface S
0pt
for this pore-throat volume is
computed by combining the surface area of the original rock, S
0r
(,
k) obtained from Eq. 5, with the surface area of the deposited solids
and polymers. The rock porosity is corrected for deposited material

pt
. Then, k
pt
(
pt
, S
0pt
) is calculated by Eq. 5, and finally the new
effective permeability is obtained by harmonic averaging with k:



pt
p s
pt
r s s
a
S
S a S S

+ j
(
,
\
,
(

( ) + +
1
1
0
0 0 0
,
pp p
s p
pt pt pt
a
k
a
k S

( )
( ) + +
( )

( )
+
1
0
,
,
eff
11
1
1

,
,
]
]
]
]

j
(
,
\
,
(

a
k
a f
pt
s p
ic
,
max ,

. . . . . . . . . . . . . . . . (40)
The proposed methods have some interesting differences from
common phenomenological models that relate the permeability reduc-
tion only to the amount of deposited material or the reduction in poros-
ity [e.g., k
eff
/k
i
(/
i
)
m
]. With the indicated model, any dependency
on the nature of the deposited particles and on the initial permeability
must be set explicitly by changing the model exponent m. By using
the Carman-Kozeny approach, properties of the trapped particles are
automatically included through their specific surface area. As a result,
the same amount of trapped material will result in larger permeabil-
ity reduction when the particles are smaller. The trapping rate will
increase at lower permeability because of a smaller pore diameter D

,
but the relative reduction in permeability per amount of trapped mate-
rial will be reduced. This is a result of decreased difference between
the core permeability and the internal-cake permeability.
A presumption in the permeability-reduction model is that all
streamlines see the same amount of trapped material. This require-
ment applies to the computational volume enclosed in a single grid
cell. With significant heterogeneities on a scale smaller than the grid
cell (i.e., if grid cells are populated with average properties), then the
presented model will be less reliable. Furthermore, the accuracy in
computed permeability reduction will not be better than the trapping
model used (Eq. 36). One obvious simplification is that we disregard
the pore-size distribution, which may result in underestimation of
initial deposition rate when the first narrow pathways are plugged.
However, significant impairment will not occur until the main flow
paths start to plug, and, presumably, this part of the process is suf-
ficiently represented by using average properties.
Pore-Lining Retention. The pore-lining retention concerns
material attached to the pore surfaces. The attachment of particles
to the rock surface is governed by a balance between attractive
and repulsive forces. This is modeled with a Langmuir-type kinetic
equation for adsorption A:
d
d
A
t
c Q A A
k
k k m k k k k

( )

3 4 , , ,
, k s, p. . . . . . . . . . . . . . . (41)
Q
m
is the adsorption capacity, c the volumetric concentration, and

3
and
4
are rate parameters representing adsorption and desorp-
tion, respectively. The same equation is used for both polymers and
solids. For a mixture with several solid and polymer components,
the following apply:
One set of polymer-adsorption parameters is used for all poly-
mers. The maximum adsorption capacity for individual polymer
components is according to their relative concentration and sums
up to Q
m,p
. Adsorption between polymers is competitive.
One set of solid-adsorption parameters is used for all solids.
The maximum adsorption capacity for individual solid components
is according to their relative concentration and sums up to Q
m,s
.
Adsorption between solids is competitive.
The model assumes that polymer and solid adsorption can
be decoupled (i.e., that no competitive adsorption between solid
and polymer exists).
A main reason for repulsive forces is the electrical charge on
rock and particle surfaces. An increasing electrolyte concentration
will reduce electrostatic repulsion. Also, a particle may depart
from the rock surface as a result of a high viscous drag. These
mechanisms are included in the desorption parameter
4
through
the pressure gradient p and the effective salinity C
se
terms:

4 40
41
6
1
1
10
41
, ,
,
,
k k
k
p
k
+

j
(
,
\
,
(
,

,
,
]
]
]
]

++C
se
k

42,
, k s, p. . . . . . . . . (42)
The last salinity term is optional and requires the presence of at
least one cation in the simulations. C
se
is defined as the equivalent
concentration of a reference cation, normally sodium. p in Eq.
42 is corrected for the effect of pore-throat plugging (i.e., this is
subtracted from the actual pressure gradient).
Effective Salinity. Both the state of swelling clays and the attach-
ment of nes to the pore surface are strongly dependent on the type
September 2010 SPE Journal 757
of exchangeable cations, such as K
+
, Na
+
, Mg
2+
, and Ca
2+
, and on
the total salinity. The equilibrium between cations in solution and
those attached to the clay surface is approximated by the mass-
action equations (Hirasaki 1982; Hirasaki and Lawson 1986).
f
f
K
X
X
Na
z
X
XC
Na
z
i
i
i
i
i
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (43)
where X
i
is the bulk phase concentration of Cation i in N (normality
or mEq/mL), z
i
is the valence, and f
Xi
is the fraction of Cation i at
the clay surface. K
XCi
is an ion-exchange constant for each cation.
X
Na
and f
Na
represent the bulk phase concentration and the surface
fraction, respectively, of the reference sodium (Na
+
) cation.
To evaluate the salinity effect in mixed brines on, for example,
clay stability, we adapt the effective salinity model from UTCHEM
(Pope and Nelson 1978; Hirasaki 1982) and extend it to any num-
ber of cations. Na
+
is used as a reference cation, and the effective
strength of other cations is transformed to an equivalent amount
of Na
+
. The effective salinity, C
se
, is calculated from the fractional
concentrations at the clay surface (for n cations):
C
X
f
se
Cl
ci Xi
i
n

1
1
1

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (44)
The total salinity is here represented by X
Cl
(aqueous chlorine
concentration), which is equal to the sum of all cations in solution.

ci
is the effective salinity parameter for Cation i.
Clay Swelling. The net transport of water into or out of a swelling
clay is driven by osmotic forces, a gradient in the effective salinity.
The swelling is modeled as a kinetic water-adsorption model with
C
se
as the driving force in both terms:
d
d


sw sw
se cse
sw sw sw se sw
t
k
C
Q k C
+
( )
1
2

max
, . . . . . . . . . . . . (45)
where
sw
is the swelling state defined as the volume ratio of adsorbed
water to clay; k
sw1
and k
sw2
are adsorption and desorption rate param-
eters, respectively, and Q
swmax
is the maximum swelling state.
The effect of the swelling is that the pore volume available
for flow will be reduced. We model this by reducing the effective
water-phase saturation by an amount equal to the trapped water.
Consequently, the relative permeability to water will decrease
during swelling. The oil relative permeability will be affected
indirectly through a decrease in maximum oil saturation equal to
trapped water. The amount of trapped water is corrected for the
initial content of water in the clay.
Results
This section presents simulations of various static and dynamic
mud-filtration experiments and a dilute particle-transport experi-
ment, and demonstrates simulations of fines migration and clay
swelling. Finally, laboratory results are used in simulation of a
North Sea well. Mud compositions are given in Table 1.
Input parameters used in the simulations (if not mentioned in
the text) are given in Table 2 (polymers), Table 3 (solids), Table 4
(various model parameters), Table 5 (pore-throat trapping), Table 6
(pore-lining adsorption), Table 7 (relative permeability), and Table 8
(dynamic filtration).
TABLE 1MUD COMPOSITIONS (g/L)
Component Mud 1 Mud 2 Mud 3a Mud 3b

Polymer (P1, P2) 25 25
Xanthan 4 4
Starch 20 20 20
CaCO
3
100 100 50 50
Drilling
solids (OCMA)
40
Barite 155
TABLE 2POLYMER PROPERTIES*

P1 P2 Xanthan Starch
Viscosity, Low Shear

[ ] (mL/g) 1500 3200 5500 600
k' 0.4 0.4 0.4 0.4
k'' 0.01 0.01 0.07 0.07
M
w
exponent, a 0.8 0.8 0.8 0.5
B
vT
[1/C] 0 0 0.006 0.001
T
0
(C) 20 20 20 20
Shear Thinning

P

1.7 1.7 1.8 1.5

hf
(s
1
) 2 2 1.5 5
Size Parameters

M
w
(g/mol) 10
7
10
7
3 10
6
5 10
6
ln(M
w
) 1 1 2 2
D
ch
( m) 0.0008 0.0008 0.001 0.0012
(g/cm ) 1 1 1 1
Computed Properties

S
0
(m /cm ) 5000 5000 4000 3333
Mean D
hp
( m) 0.414 0.498 0.428 0.242
Filter Cake Properties

o
0.85 0.85 0.85 0.85
n
c
(bar
1
) 0.095 0.095 0.08 0.08
* Numbers in italic do not influence static fluid loss computation.
758 September 2010 SPE Journal
Static Filtration. The lter-cake model is demonstrated by computing
static high-pressure/high-temperature (HP/HT) ltration experiments
with two muds performed at 85C and 850 psi on 20- and 60-m
ceramic disks. The solids size distributions for the two mud systems
are given in Fig. 2. The manufacturer permeability of the two disks
is 5 and 20 darcy, respectively. These permeabilities and a porosity of
0.37 were used in the calculations. Filtration volumes are computed
for a 45-cm ltration area [American Petroleum Institute (API)].
The 0.0722 volumetric concentration of solids in Mud 1 (S1 in
Table 3) includes a mixture of sized CaCO
3
and barite. Average poly-
mer properties are used arbitrarily to match the filtration volume; see
P1 in Table 2. The polymer concentration was set to 0.025.
Mud 2 contained 0.0377 volume-fraction-sized CaCO
3
(S2 in
Table 3). The only other parameter changed in the Mud 2 calcula-
tions was the intrinsic polymer viscosity [], which was increased
from 1500 to 3200 mL/g (P2 in Table 2). All other parameters
remained unchanged.
Details from computation of the spurt-loss and filtration volumes
for Mud 1 on the 5-darcy filter are given in Table 9. The main steps
include
1. Determine critical size of solids that can enter the formation
(
s
D

0.435.8 14.3 m) (Eqs. 6 and 22).


2. Use size distribution to find concentration and specific sur-
face area of particles trapped at the disk surface. Compute average
deposition thickness needed to form the initial cake, h
sp
n
sp
6/S
0sp

21.1 m (Eq. 23).
3. Compute spurt-loss volume, V
sp
1.25 mL (Eq. 24).
4. Determine average filter-cake permeability k
ca
9.310
6
md
and <
c
> 0.416 from average cake properties, <
o
> 0.515, <n
c
>
0.0244 bar
1
, and p 58.6 bar (Eqs. 11, 12, and 26).
5. Finally, compute the static filtration parameter C
st
0.0309
mL/cmmin
0.5
and fluid loss V
f
(30 minutes) 8.86 mL (Eq. 4).
Experimental and simulated filtrate volumes are plotted against
the square root of time in Fig. 3 (Mud 1) and Fig. 4 (Mud 2). We
TABLE 3SOLIDS PROPERTIES (S
0
COMPUTED)

S1*

S2*

CaCO
3
*
Mud 3
OCMA
Mud 3b
Fe
2
O
3
*

D
p
( m)**

18.44 29.4 7.05 1.5
0.856

St Dev

1.4 1.34 0.8 1
0.95

sh
1.8 1.35 1.25 1.5
1.8
S
0
(m /cm ) 1.5 0.67 1.5 9.91
19.6

(g/cm ) 2.6 2.6 2.7 2.7
5.4
Filtercake

o 0.4 0.4 0.4 0.4


0.4
n
c
(bar
1
)
0.0001 0.0001 0.0001 0.001
0.0005

* Tabular distribution, mean Dp, and standard deviation are computed.
** D
p
corresponding to mean ln(D
p
).

Standard deviation in ln(D


p
).
TABLE 4VARIOUS MODEL PARAMETERS,
VALUES USED IN SIMULATIONS
Variable Value Remarks
s
0.4 size factor in spurt loss
p
0.1 size factor for polymers
n
sp
1 spurt-loss factor
3 tortuosity factor
TABLE 5PORE-THROAT-TRAPPING PARAMETERS

Solids Polymers
t
(fraction) 0.035 0.05
10
(cm
1
) 40 1.5
11
2 2.5
20
(cm
1
) 0.0001 0.0015
22
(bar/cm) 0.01
22
0.5
f
pm
1
f
pt
0.05
f
pt2
0.1
TABLE 6PORE-LINING-ADSORPTION PARAMETERS
Solids Polymers Fe
2
O
3
Fines
Q
m
(pv) 0.05 0.001 0.05 0.2
30
(min
1
) 0.1 0.1 0.1 0.03
31
0 0 0 0
40
0.0001 0.0002 0.0001 0.0001
41
(bar/cm) 10 0.1 10 10
41
3 1 3 0.25
42
1.5 1.5 1.5 1.5
RPR
max
2 2 2 2
200 1000 200 1000
TABLE 7RELATIVE PERMEABILITY,
COREY PARAMETERS
Simulation k
rwe
E
w
S
rw
k
roe
E
o
S
or

FD1-FD3 0.15 3 0.2 0.77 1.5 0.2
FD15 0.15 3 0.2 0.65 1.5 0.2
B14 0.1 4 0.185 0.75 3 0.185
Field 0.4 3 0.14 0.65 4 0.1
TABLE 8DYNAMIC-FILTRATION PARAMETERS
Simulation f
n
n
F

e
(s
1
)
y
(Pa)
DS2 0.1 1.25 0 5000
DS3 0.1 0.5 67 5000
DS4 0.05 0.8 1 1000
Mud98067 300 0.5 1 10
0
0.2
0.4
0.6
0.8
1
0.1 1 10 100 1000
Particle Diameter, m
C
u
m
u
l
a
t
i
v
e

D
i
s
t
r
i
b
u
t
i
o
n
,

F
,

v
o
l
.

f
r
a
c
t
i
o
n
Mud 1
Mud 2
Mud 3
Fig. 2Particle-size distribution.
September 2010 SPE Journal 759
observe that increasing the permeability (pore size) of the disks
results in a higher V
sp
. This is because a smaller fraction of the
solids will be held back at the disk surface and, in addition, the
average size of retained particles will be larger so that a larger
volume is required to form the initial cake. V
sp
is more than doubled
for Mud 2 compared to Mud 1. This is mainly because of the lower
solid concentration. All these differences are well reproduced in the
calculated curves, as clearly illustrated in Figs. 3 and 4.
The second observation is that the filtration curves are parallel
to each other (i.e., the filtration parameter C
st
does not depend on
the properties of the formation but only on the mud properties and
the differential pressure over the cake).
Fig. 5 shows computed filtrate volumes with Mud 1 for a wide
range of permeabilities. We show both the analytically computed V
f

and V
sp
, ignoring the pressure drop over the disk, and the simulated
V
f
(30 minutes) using a 10-cm core. We observe that V
sp
starts to
increase above 10 md, while the net filtration volume after the
spurt loss (V
f
V
sp
) is constant in the analytical calculations. The
simulated V
f
deviates from the computed value at low permeability
(0.01 md) when the resistance over the core becomes significant
compared to the filter-cake resistance. Also, at very high perme-
ability (100 darcies), the simulated V
f
is lower than the computed
V
f
because of internal blockage of pores that reduces the effective
pore size of the formation.
Blaxter Sandstone. The results from three static ltration experi-
ments using Blaxter sandstone are given in Table 10. The core
dimensions were 8-cm length and 2-in. diameter, with an effective
ltration area of 16.8 cm
2
. The permeability was approximately
150 md, and the porosity was 20%. The mud system used in Core
Experiment FD1 contained 50 g/L sized CaCO
3
with size distri-
bution given in Fig. 2 (Mud 3a), 4 g/L xanthan, and 20 g/L low
reticulated starch. 40 g/L drilling solids [Oil Company Materials
Association (OCMA) clay] was added in Experiments FD2 and
FD3 (Mud 3). All the coreoods were conducted at 90C. An
overbalance pressure of 15 bar was used in FD1 and FD2, and 35
TABLE 9COMPUTATION OF STATIC FILTRATION, MUD 1, A
c
= 45 cm
Model parameters
s
= 0.4
p
= 0.1

n
sp
= 1

[ ] = 1,500 mL/g
s
= 0.339 cp

Disc properties
K (md) = 5000 = 0.37 Ac = 45 cm
2
D

= 35.8 m
s
D

= 14.3 m

Spurt loss Fraction >
s
D

= 0.589 S
0sp
= 0.284 m
2
/cm
3
h
sp
= 21.2 m C
1sp
= 0.0763 V
sp
= 1.25 mL

Mud

Concentration
(volume fraction)
Effective Concentration
(volume fraction) S
0
(m /mL)
o n
c
(bar
1
)
Polymer 0.0250 0.0249 5000 0.850 0.0950
Solids 0.0722 0.0722 1.5 0.400 0.0001
Total 0.0972 0.0971 1283 0.515 0.0244
Filtration p = 58.6 bar
<
c
> k
ca
(md) C
1
C
st
(mL/cm min
0.5
)

V
f
30 minutes (mL)
0.416 9.27E-06 0.1993 0.0309 8.86
0
2
4
6
8
10
12
0 1 2 3 4 5
Square Root of Time, minutes
V
f

,

m
L
6
20 m 1 60 m 1
20 m sim 60 m sim
Fig. 3HP/HT filtration at 850 psi and 85C, experimental and
simulated for 20- and 60-m filter disks, Mud 1. (A
c
= 45 cm).
0
2
4
6
8
10
12
14
16
0 1 2 3 4 5 6
20 m 3 60 m 3
20 m sim 60 m sim
Square Root of Time, minutes
V
f

,

m
L
Fig. 4HP/HT filtration at 850 psi and 85C, experimental and
simulated for 20- and 60-m filter disks, Mud 2. (A
c
= 45 cm).
0.1
1
10
100
0.001 0.1 10 1000 100000
K, md
F
i
l
t
r
a
t
e

V
o
l
u
m
e
,

m
L
V
sp
V
f
V
sp
V
f
V
f sim
Fig. 5Effect of the porous medium on filtrate volumes. Ana-
lytical spurt loss (V
sp
) and filtrate volume (V
f
) and simulated V
f

(30 minutes) with a 10-cm core (V
f-sim
). (A
c
= 45 cm), Mud 1.
760 September 2010 SPE Journal
bar in FD3. API ltration (100 psi and 25C) and HP/HT (500 psi
and 90C) are also reported in Table 10.
Properties of the polymers and solids are given in Tables 2 and
3, respectively. Fig. 6 shows a good agreement between measured
and computed shear-thinning mud viscosity. Other model param-
eters used in the simulations are given in Tables 4, 5, and 6.
Filtration. All the experimental fluid-loss volumes in Table 10
are well reproduced in the simulations. This indicates that the
effects of pressure, temperature, and composition are all well taken
care of in the simulations. The pressure sensitivity (735 bar) is
handled by the compressible-filter-cake model. The simulated
effect of temperature on filtrate volume is only through the change
in the brine viscosity, from 0.926 cp in the API test (25C) down
to 0.318 cp in the core tests (90C). The effect of composition is
more complex. In the core experiments, we observe a reduction
in V
f
when OCMA clay is added (FD1 vs. FD2). In the calcula-
tions, the filter cake in FD2 will be thicker but with a higher k
ca

because of the reduced relative polymer content. However, adding
OCMA clay decreases the average solids particle size, which may
potentially trap more polymer in the cake. In the simulations,
approximately 9% of the xanthan and 30% of the starch is passing
through the filter cake in FD1. In FD2, these numbers are reduced
to zero for xanthan and 1% starch. The increased entrapment of
polymer in the filter cake will reduce k
ca
, and this is the reason, at
least in the simulations, for the observed decrease in V
f
between
FD1 (12 mL) and FD2 (10 mL).
The relative viscosity of the filtrate produced from Mud 3a was
measured to be approximately 6 and showed shear-thinning proper-
ties. This verifies that a significant amount of polymer was passing
through the filter cake in FD1. The lower level of simulated poly-
mer migration through the cake in FD2 agrees well with measure-
ments taken by Snchez et al. (2004), where 12% of the polymer
was measured (as total organic carbon) to pass through the cake
for a mud system similar to Mud 3b. However, the actual values of
the model parameters used to simulate this phenomenon are highly
uncertain. In the simulation, we identify the degree of polymer
entrapment in the cake as a potentially important variable for fil-
tration rates; however, the properties used in these calculations are
poorly known. The polymer molecular weight M
w
is well known
as a physical property, but its actual value and its size distribution
are highly uncertain. The computation of molecular hydraulic size
from [] and the trapping parameter
p
are also uncertain. Finally,
the OCMA properties used are also approximate.
Return Permeability. The oil return permeabilities in Table 10
are obtained after backflooding the core with oil at increasing rates:
1, 2.5, 5, and 8 mL/min, for 30 minutes at each rate. Simulated and
experimental return permeabilities are in the same range, 8090%
of initial oil permeability k
oi
. The simulated differential pressure
over the core and the development of the return permeability RKO
during the backflood of FD1 are shown in Fig. 7. We see that RKO
is highly dependent on the length of the backflooding period. The
profile is governed by relative permeabilities, which are unknown,
and viscosity ratio of oil to water, which may be reduced because
of polymer invasion. The relative permeabilities used (see Table 7)
are typical for water-wet medium, with a final tuning to the results
in FD1, except for the oil endpoint permeability.
The average S
w
shown in Fig. 7 is still well beyond its initial
value of 0.2. Also shown is RKOE, an estimated return permeabil-
ity at maximum oil saturation. If the flood were continued, RKO
(0.824) should finally reach RKOE (0.916). Note that capillary
pressure is ignored (zero) in these simulations and that additional
uncertainty in return permeabilities may be caused by capillary
end effects.
Damage Profile. Another issue much discussed is what the
damage profile looks like in the invaded zones. The simulated
damage profile (RKO) for FD1 is shown in Fig. 8 together with
profiles of trapped CaCO
3
and polymers. CaCO
3
particles are
trapped in the first millimeters of the core, while polymer shows
a deep penetration. It seems that trapped polymer has reached a
plateau extending approximately half the filtrate-invasion volume.
The damage profile (RKO) follows the trapped-polymer profile.
The amount of external CaCO
3
particles introduced during the
spurt-loss period is obviously very small, and there is no realistic
way to simulate any serious damage from trapped CaCO
3
particles
alone. The observed damage, therefore, is caused most likely by
the polymer.
Another experiment, FD8, shows very serious plugging when
only the polymer mixture is injected (4 g/L xanthan plus 20 g/L
starch), shown in Fig. 9. When simulating FD8 with polymer-trap-
ping parameters given in Table 5, only the first part of the pressure-
drop curve was matched. The pressure increase during long-term
TABLE 10EXPERIMENTAL AND SIMULATED
FLUID-LOSS VOLUMES WITH MUD 3
FD1 FD2 FD3
p (bar) 15 15 35
Drill solids (g/L) 0 40 40
API (100 psi, 25C) exp 6 5
V
f
30 minutes (mL) sim 5.9 4.7
HP/HT* (500 psi, 90C) exp 12 12
V
f
30 minutes (mL) sim 12.4 10.6
Core** (90C) exp 12 10 11
V
f
240 minutes (mL) sim 12.1 10.3 11.1
Oil return exp 83 82 86
Permeability (%) sim 82.9 86.8 85.4
* HP/HT filtration volumes converted to API filtration area.
** Core filtration area is 16.8 cm
2
.
1
10
100
1000
10000
1 10 100 1,000 10,000
Shear Rate, 1/sec
V
i
s
c
o
s
i
t
y
,

m
P
a

s
Calc
Exp.
Mud 3a
Fig. 6Mud 3a, experimental (converted Fann viscosities) and
calculated viscosity, 25C.
0
0.15
0.3
0.45
0.6
0.75
0 20 40 60 80 100 120
Time, minutes
P
r
e
s
s
u
r
e

D
r
o
p
,

b
a
r
0
0.2
0.4
0.6
0.8
1
K
o

/
K
o
i

a
n
d

<
S
w
>
dp RKO
RKOE <S
w
>
FD1
1 mL/min
8 mL/min
5 mL/min
2.5 mL/min
Fig. 7Simulated back production, FD1. Differential pressure
(dp), current (RKO) and maximum (RKOE) oil return perme-
ability, and average S
w
.
September 2010 SPE Journal 761
injection shown in Fig. 9 was approximately simulated using a
polymer modification factor f
pm
0.15 (see comment after Eq. 39).
The amount of polymer that will invade the core in typical mud-fil-
tration experiments is well within the part that is matched without
modification of the polymer effect (f
pm
1, as in Table 5).
In Fig. 10, the backflooding of Core B14 is extended with an
extra step where the final oil in place is displaced with a highly
viscous oil (M82) at a reduced rate. Using the pressure derivative
with respect to the front position of M82, the permeability profile
along the core can be estimated. We approximate the front position
by assuming constant S
o
throughout the core. In Fig. 11, we show
the estimated k
o
profile (Experiment VF) together with k
oi
. We also
show profiles from simulation of the experiment and an estimated
k
o
profile from the simulated pressure history (Simulation VF).
Note that the estimated increase in k
o
toward the outlet end in the
oilflood, which is to the left in Fig. 11 and the mud side, is a result
of dispersion of the viscous-oil front. The same behavior was seen
when applying the method on simulated data (Simulation VF).
Core B14 has a length of 4.7 cm and was subjected to static mud
exposure for 50 hours (Mud 3b), during which close to 2 PV of
filtrate had entered the core. The results in Fig. 9 strongly indicate
that k
o
is reduced throughout the whole core, which supports the
previous results (Fig. 8), indicating that polymers may cause deep
and rather evenly distributed damage.
Dynamic Filtration. Two examples of simulating dynamic experi-
ments are given in Figs. 12 through 14. The rst example is Core
Test FD15 (Blaxter sandstone) with Mud 3b (described earlier).
The experiment is simulated with different sets of parameters (f
n
,
n
F
, ,
y
), given in Table 8. The plot of V
f
vs. time in Fig. 12 has
0
0.2
0.4
0.6
0.8
1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
0
0.0002
0.0004
0.0006
0.0008
0.001
T
r
a
p
p
e
d

C
a
C
O
3

1

a
n
d

p
o
l
y
m
e
r

1
0
0
,

P
V
RKOE
RKO
CaCO
3
Xanthan
Starch
K
o

/
K
o
i

Fig. 8Simulated oil permeability reduction, current (RKO) and estimated final (RKOE), and pore-throat-trapped solids (CaCO
3
)
and polymers.
0
5
10
15
20
25
30
35
0 1 2 3 4
PV Injected
P
r
e
s
s
u
r
e

D
r
o
p
,

b
a
r
5
Exp. FD8
Sim. f
pm
=1
Sim. f
pm
=0.15
Fig. 9Xanthan+starch injection, 150-md Blaxter sandstone.
0
0.1
0.2
0.3
0.4
0.5
0.6
0 5 10 15 20 25 30 35 40
PV Injected
P
r
e
s
s
u
r
e

D
r
o
p
,

b
a
r
0
0.1
0.2
0.3
0.4
0.5
0.6
O
i
l

R
e
t
u
r
n

P
e
r
m
e
a
b
i
l
i
t
y
,

R
K
O
Sim.
Exp. CLS370
Exp. M82
RKO
B14: back production
Fig. 10Back production with low-viscosity oil (CLS370) and
highly viscous oil (M82) in Experiment B14.
0
20
40
60
80
100
120
140
160
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Core Position, x/L
K
o
Sim. Sim VF
Exp. VF K
oi
K
o max
B14
Fig. 11Interpretation of oil permeability profile (Experiment VF) from viscous-oil flood, B14.
762 September 2010 SPE Journal
a nonuniform development of its slope, which may suggest some
experimental problems. This is seen more clearly when plotting
the slowness (inverse velocity) vs. V
f
, as in Fig. 13. Simulation
DS3 puts more weight on the last part and assumes that a steady-
state limit of approximately 6000 s/cm is reached. DS4 puts more
weight on the rst part, while DS2 is in between. DS2 and DS4
have not reached steady state within the 4-hour time frame. In DS2,
no erosion at all is assumed, only a reduced attachment rate. This
illustrates the problem of upscaling a short dynamic experiment to
the eld situation, which will be shown in a later section.
The second dynamic case (in Fig. 14) is a filtration experiment
taken from Dewan and Chenevert (2001). A bentonite mud with
solid content c
sm
0.18 is used. Appropriate rheologic mud prop-
erties were simulated using a polymer component. The solid-size
distribution was taken from one of the other muds (Duratherm)
used in their paper. The simulated slowness curve in Fig. 14
matches the experimental curve very well. The experiment starts
with dynamic filtration (100 s
1
) at 500 psi, switches to static con-
ditions, switches back to dynamic conditions again, and ends with
increasing the pressure to 1,500 psi. Each period was 30 minutes,
which was insufficient to reach steady-state conditions. After the
pressure increase, we first see a short transient period where the
compressible cake adapts to the new pressure. Then, the curve will
slowly progress toward the same steady-state limit as before the
pressure increase, although at a slower rate because more mass
must accumulate in the cake at the higher pressure.
The uncertainties in the particle-size distribution and the rheologic
properties make the value of the model parameters in Table 8 less gen-
eral. The main purpose, however, was to demonstrate the capability of
the dynamic model to reproduce typical experimental behavior.
Particle Retention. An experiment (Experiment 1) taken from Al-
Abduwani et al. (2003) is used to demonstrate the pore-throat-trap-
ping model. In this experiment, a dilute 20-ppm hematite (Fe
2
O
3
)
particle suspension was injected into a Bentheimer sandstone core.
The particle size was stated to be in the range of 0.1 to 5 m, with
65% less than 1 m. The size distribution used in our simulations
is given in Fig. 15, and additional information is given in Table 3.
Core data used in the simulations are L 12.2 cm, k 1300 md,
0.22, and S
w
1. The trapping parameters for solids tuned to
this experiment (Table 5) are also used in the other simulations
presented in this paper.
By using dx/dt u
w
/(S
w
) in Eq. 36, setting
2
to zero, and integrat-
ing, we find the effluent concentration for a given particle size to be
c c L ( )
0 1
exp , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (46)
0
5
10
15
20
25
30
35
40
45
50
0 50 100 150 200 250
Time, minutes
F
i
l
t
r
a
t
e

V
o
l
u
m
e
,

m
L
exp.
DS2
DS3
DS4
FD15
Fig. 12Dynamic filtration with Blaxter sandstone, Mud 3b.
Simulation of Experiment FD15.
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
0 10 20 30 40 5
V
f
, mL
S
l
o
w
n
e
s
s
,

s
/
c
m
0
exp.
DS2
DS3
DS4
FD15
Fig. 13Slowness vs. V
f
in dynamic-filtration experiment FD15.
0
10000
20000
30000
40000
50000
60000
0 30 60 90 120
Time, minutes
S
l
o
w
n
e
s
s
,

s
/
c
m
Exp. Mud 98067
Sim
500 psi
100 seconds
1
500 psi
0 seconds
1
500 psi
100 seconds
1
1500 psi
100 seconds
1
Fig. 14 Experimental and simulated dynamic filtration. Experi-
mental data are replotted from Dewan and Chenevert (2001).
0
0.2
0.4
0.6
0.8
1
0.1 1 10
Particle Diameter, m
C
u
m
.

V
o
l
.

F
r
a
c
t
i
o
n
0
0.2
0.4
0.6
0.8
1
R
e
l
a
t
i
v
e

E
f
f
l
u
e
n
t

C
o
n
c
.
F Hematite
C
eff
C
eff
(Dlim)
Fig. 15Hematite particle distribution used in simulation. Relative effluent concentration is for the Bentheimer core described
in the text, computed with (Dlim) and without a lower size limit for trapping (
ts
).
September 2010 SPE Journal 763
where
1
(D
p
) is given by Eq. 37. The computed size-dependent
relative effluent concentration for the present core is shown in the
size-distribution plot (Fig. 15). The average effluent concentration
is estimated from the size distribution to be 13.1 ppm when 20.7
ppm is injected.
Simulated and experimental results are compared in Figs.
16 through 18. If pore-throat trapping is considered as the only
retention mechanism, the simulated effluent concentration seen
in Fig. 16 starts at the 13 ppm analytical value and then declines
slowly as an effect of previously trapped material. The simulation
was extended beyond the experimental time period, and, at some
point, when an external filter cake has been formed, the effluent
concentration dropped rapidly to zero. The experimental curve is
different; it breaks through at a low concentration and increases
rather slowly toward a plateau after approximately 300 PV. This
curve shape may be explained in part by a gradual change in the
trapping pattern caused by a distribution in pore-throat size, but it
is also influenced by other retention mechanisms. A closer match
to the experimental profile is obtained by including pore-lining
adsorption in the experiment.
A total of 662 PV was injected in Experiment 1. The final
distribution of hematite and permeability along the core is shown
in Figs. 17 and 18, respectively. In both cases, the simulated and
experimental profiles match. The experimental data represent aver-
ages over core segments. A corresponding segmental permeability
distribution computed from the simulation results almost overlaps
the experimental curve shown in Fig. 18.
Fig. 19 shows the specific surface area of trapped hematite
particles throughout the core from the simulation. A corresponding
effective average particle diameter, computed as D
p
6
sh
/S
0
, is also
shown. Obviously, the average size of trapped particles is larger at
the inlet end because smaller particles will penetrate deeper.
This case demonstrates the capability of the pore-throat
model to reproduce experimental data not only in terms of overall
0
2
4
6
8
10
12
14
16
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800
PV Injected
E
f
f
l
u
e
n
t

H
e
m
a
t
i
t
e
,

p
p
m
Sim
Exp.1
Exp.2
Exp.4
Sim A=0
Fig. 16Experimental and simulated effluent profile for hematite. Pore-lining adsorption is turned off in Sim A = 0. Experimental
data are replotted from Al-Abduwani et al. (2003).
0
0.001
0.002
0.003
0.004
0.005
0.006
0.007
0.008
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
F
e
O
3

C
o
n
c
e
n
t
r
a
t
i
o
n
,

P
V
Sim
Exp.
Fig. 17Experimental (Al-Abduwani et al. 2003) and simulated hematite distribution in core after 662-PV injection.
0
0.2
0.4
0.6
0.8
1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
N
o
r
m
a
l
i
z
e
d

P
e
r
m
e
a
b
i
l
i
t
y
Sim
Sim. segment
Exp. segment
Fig. 18Experimental (Al-Abduwani et al. 2003) and simulated permeability profile in core after 662-PV injection.
764 September 2010 SPE Journal
permeability reduction but also by matching internal-damage and
particle-deposition profiles. Although promising, it is still uncer-
tain how well the parameters in Table 5 will adapt to different rock
types, other permeability levels, and with other fluid/solid systems.
The test was performed with hematite particles in distilled water
flooded through a highly permeable homogeneous Bentheimer
core and with only approximate knowledge of the particle dis-
tribution. One might expect increased trapping rate and more
damage in less-uniform rocks with poorly sorted sands and higher
tortuosity. One benefit of the applied Carman-Kozeny approach is
that core properties such as permeability, porosity, and tortuosity
are included in the computations; however, additional testing is
required before one can tell if this is done adequately.
Fines Migration and Clay Swelling. Simulations of nes migra-
tion and clay swelling upon injection of low-salinity brine are dem-
onstrated in Fig. 20. In the nes simulation, 1% of the rock matrix
is dened as nes, held in place using the pore-lining-adsorption
model. In the clay-swelling simulations, 2% of the rock matrix is
dened as swelling clay with a maximum swelling capacity of 10
(Q
swmax
in Eq. 45). Other parameters in Eq. 45 are k
sw1
0.0002
mEq/(mLmin) and k
sw2
0.001 mEq/(mLmin). The effective
salinity C
se
is essentially equal to the NaCl concentration indicated
in Fig. 20, and is 0.7 mEq/mL with the initial brine.
Field-Scale Simulations. To test the upscaling to eld conditions,
we use Mud 3b on a horizontal well from the North Sea. Perme-
ability and porosity distribution along the wellbore is given in Fig.
21. We set the wellbore diameter to 0.24 m and use a constant pres-
sure boundary at 100 m in our 2D radial model. The temperature is
90C. Filtrate invasion is simulated for a period of 250 hours with
an overbalance pressure of 86 bar. Then, we simulate cleanup by
back producing the well using a 30-bar drawdown for 24 hours.
Four cases are run, FS1 with static ltration and FS2 through
FS4 with dynamic conditions at shear rate of 170 s
1
. Cases FS2
through FS4 correspond to various matches to the dynamic core
experiment FD15 and dynamic-ltration simulations DS2 through
DS4 shown in Table 8.
The simulations are summarized by main numbers in Table 11.
Typical results from the simulations are demonstrated by Simula-
tion FS4 in Figs. 22 through 25. Fig. 22 shows the saturation
distribution around the well after mud exposure for 200 hours
and after the 24-hour back-production period. The variation in
penetration depth is mainly a result of the porosity distribution
along the wellbore because filtration rate through the filter cake is
independent of the formation properties. After the back production,
we see that there is still a significant amount of water around the
well, indicating that substantial production is needed to re-establish
original fluids saturation. This is a result of bypassing caused by
the heterogeneous permeability distribution.
Figs. 23 and 24 show the computed permeability impairment
along the well. The normalized oil permeability for each zone is
computed by two methods in Fig. 23: (1) as a productivity index
0
2
4
6
8
10
12
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
S
0

T
r
a
p
p
e
d

F
e
O
3

(
m

/
c
m

)
0
0.5
1
1.5
2
2.5
3
T
r
a
p
p
e
d

P
a
r
t
i
c
l
e

D
,

m
S0
D
p
(m).
Fig. 19Profiles of specific surface area and corresponding particle size of hematite in core after 662-PV injection (simulated).
0.1
1
10
100
1000
0 5 10 15
PV Injected
P
r
e
s
s
u
r
e

D
r
o
p
,

b
a
r
Fines
Swelling clay
Brine
0.5 N
NaCl
0.1 N
NaCl
0.01 N
NaCl
0.001 N
NaCl
Brine
Fig. 20Simulation of fines mobilization and clay swelling.
0.0 0.5 1.0 1.5 2.0
0
200
400
600
800
1000
1200
1400
lg(md)
2.686
1.765
0.845
0.075
0.996
Radial Direction, m
Permeability
W
e
l
l
b
o
r
e
,

m
0.0 1.0 2.0
0
200
400
600
800
1000
1200
1400
fraction
0.248
0.196
0.145
0.093
0.041
Radial Direction, m
Porosity
W
e
l
l
b
o
r
e
,

m
Fig. 21Field case, permeability (logarithmic) and porosity
distributions.
September 2010 SPE Journal 765
computed from the rate in each zone and (2) by analytically
ignoring crossflow between zones (RKO). The error introduced
by ignoring crossflow is very small in this case and almost zero
for the total well with relative PI
oil
and RKO equal to 0.794 and
0.792, respectively, in FS4. The same results are plotted in terms
of skin distribution in Fig. 24 where S (current) is computed from
the oil rate at the end of the 24-hour back-production period and
corresponds to RKO. S (k
abs
) is computed from the reduction in
absolute permeability, corresponding approximately to RKOE. The
difference between the two skin values is because of the incom-
plete back production of water. The radial distribution of RKO and
RKOE along the well is given in Fig. 25.
TABLE 11FIELD-CASE SIMULATION RESULT,
AFTER 250 HOURS OF MUD EXPOSURE (86 bar, 170 s
1
)
AND AFTER 24 HOURS OF BACK PRODUCTION
(30-bar DRAWDOWN)
Simulation V
f
(m
3
) RKO S (current) S (absolute k)
FS1 55 0.969 0.22 0.011
FS2 504 0.872 1.0 0.005
FS2c* 225 0.924 0.6 0.008
FS3 1717 0.732 2.3 0.003
FS4 702 0.842 1.2 0.004
FS4b** 535 0.833 1.3 0.005
FS4c* 251 0.895 0.8 0.008
* Dynamic plus static periods (50 plus 25 hours); static for the last 50 hours.
** Included rate of penetration 10m/h.
0.0 1.0 2.0
0
200
400
600
800
1000
1200
1400
So
0.860
0.691
0.521
0.352
0.182
So, 200 hours #FS4
0.0 1.0 2.0
0
200
400
600
800
1000
1200
1400
So
0.860
0.690
0.520
0.350
0.180
So, 24 hours Back
production. #FS4
Radial Direction, m
W
e
l
l
b
o
r
e
,

m
Radial Direction, m
W
e
l
l
b
o
r
e
,

m
Fig. 22Oil distribution around well after 200 hours of mud
exposure at 86 bar, 170 s
1
(left) and after 24 hours of back
production at 30-bar drawdown (right).
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 200 400 600 800 1000 1200 1400
Wellbore, m
P
I
,

r
e
l
a
t
i
v
e
Oil RKO #FS4
Fig. 23Simulated FS4. Relative productivity index for oil, from simulated well rate and computed ignoring crossflow (RKO).
After 24 hours of production at 30-bar drawdown.
0.001
0.01
0.1
1
10
100
0 200 400 600 800 1000 1200 1400
Wellbore, m
S
k
i
n
S (current) S (K
abs
) #FS4
Fig. 24Simulated FS4. Skin distribution computed from current productivity and from current absolute permeability. After 24
hours of production at 30-bar drawdown.
766 September 2010 SPE Journal
From the upscaling of laboratory results with Mud 3b, we
conclude that no significant real damage to the formation should
be expected from externally introduced particles in the present
case. Other internal damage processes are not evaluated in this
example. The real damage, as indicated by RKOE in Fig. 25, is in
line with the observed polymer damage observed in the laboratory
experiments. However, one should expect temporarily reduced
productivity because of incomplete displacement of mud filtrate.
The size and duration of this decline depend on filtrate penetration
depth, relative permeability, and the variance of the permeability
distribution.
The fluid-loss volumes in Table 11 are very high; in fact, they
are an order of magnitude larger than expected. The worst case is
FS3, in which a high erosion rate is assumed. Using a steady-state
slowness of 5628 s/cm corresponding to that obtained for DS3
in Fig. 13 but corrected for the different shear rate in FS3 and
the filtration area of the well (D
w
L
w
1094 m), we compute a
filtrate volume of 1750 m compared to 1717 m given in Table
11. Similarly, the static fluid loss for the well is estimated to be
55.6 m using the filtration parameter C
st
from FD3 (0.0032 m/
mh
0.5
), which compares well to the simulated 55 m in Table 11.
The deviations are because of differences in overbalance pressure
and low-permeability zones offering flow resistance comparable
to the mudcake resistance. These computations indicate that the
field-scale simulations are correct in terms of upscaling the labo-
ratory results, but our description of the field conditions may be
insufficient. Of course, our interpretation of the laboratory results
is very important, and, in the case of FD15, the estimated dynamic
parameters are very uncertain. But all the interpretations used pre-
dicted large fluid-loss volumes in the field case. Other explanations
should be sought.
One obvious note is that the formation will be gradually
exposed to the mud as the well is drilled. In the simulations, the
total well was exposed from time zero. Exposing the formation
according to a well penetration rate of 10 m/h is shown in Fig. 26
(FS4b). However, this will result only in a moderate reduction of
the filtrate volume from 702 to 535 m. Simulating with periods
of 50 hours dynamic plus 25 hours static and, last, 50 hours of
static reduces the loss to 251 m (FS4c). In Simulation FS2, only
reduced mudcake growth and no erosion are assumed. Apply-
ing the same periodic dynamic/static history on FS2 reduces the
fluid loss from 504 to 225 m (FS2c). Variations in shear rate are
obviously important, and knowledge about how the mudcake will
respond to increased shear stress after a static period is vital for
good predictions at the field scale. Other possibilities that may
contribute to higher than expected computed fluid loss at the
field scale are that the effective shear rate is lower than projected
because of off-centered drillstring and the presence of significant
surface roughness in the well compared to the flat core ends used
in laboratory experiments.
Summary and Conclusions
Maximize is a program for computation of formation damage
imposed on the formation during drilling. Among the main physical
phenomena modeled in the program are filter-cake buildup under
static and dynamic conditions, fluid loss to the formation, transport
of solids and polymers in the formation including effects of pore-
lining retention and pore-throat plugging, and salinity effects on
fines stability and clay swelling. The models handle multicompo-
nent water-based-mud systems at both core and field scales.
Our filter-cake model differs from previous models in that we
compute the filter cakes porosity and permeability from proper-
ties ascribed to the individual solid and polymer components in
the mud. The necessary data for each component are porosity and
compressibility, if packed as a sandpack, and specific surface
area. The specific surface area is obtained from the size distribu-
tion for solid particles and from the average chain diameter of
polymers. The models capability to reproduce various fluid-loss
experiments is demonstrated. For a CaCO
3
mud, the effects of
varying temperature, differential pressure, and addition of drilling
solids are all simulated with a single set of input parameters.
The particle retention is split into two main mechanismspore-
throat trapping and pore-lining retention. We focus mainly on the
pore-throat trapping, which is the process with the main damage
potential. A dilute-particle-injection experiment was selected for
testing the model. The simulation provided excellent matching of
0.0 0.5 1.0 1.5 2.0
0
200
400
600
800
1000
1200
1400
RKO
1.000
0.750
0.500
0.250
0.000
Radial Direction, m
RKO, #FS4, 24 hours
back production
W
e
l
l
b
o
r
e
,

m
0.0 0.5 1.0 1.5 2.0
0
200
400
600
800
1000
1200
1400
RKOE
1.000
0.950
0.900
0.850
0.800
Radial Direction, m
RKOE, #FS4, 24 hours
back production
W
e
l
l
b
o
r
e
,

m
Fig. 25Normalized oil permeability after 24 hours of produc-
tion RKO (left) and RKOE computed at maximum S
o
(right),
simulated FS4.
Fig. 26Simulated FS4b, with well penetration rate of 10 m/h.
0.0 0.5 1.0 1.5 2.0
0
200
400
600
800
1000
1200
1400
So
0.860
0.696
0.532
0.369
0.205
Radial Direction, m
So, 125 hours, Sim. SF4b
W
e
l
l
b
o
r
e
,

m
September 2010 SPE Journal 767
final experimental profiles of retarded particles and permeability
along the core.
Upscaling of laboratory results to the field scale is demonstrated
for a North Sea well. Only damage from externally introduced mud
particles and polymers was considered. The simulations indicated
essentially no permanent damage, in line with the moderate poly-
mer damage observed in the laboratory experiments. However, a
temporary decline in productivity because of slow displacement of
invaded fluids was observed. The slow restoration of initial satura-
tion around the well is attributed to the heterogeneous permeability
distribution and to large fluid-loss volumes in the simulations. The
simulated fluid-loss volume is an order of magnitude larger than
that normally observed in the studied field cases. The dynamic
fluid loss had been tuned to a dynamic laboratory experiment, and
analytical calculations verify that the simulated results are correct
in terms of upscaling the fluid-loss rate at the laboratory scale.
Nomenclature
a M
w
exponent for polymer viscosity
A area, L
2
A adsorbed material (PV), solids or polymer
B
p
viscosity parameter for pressure, Lt
2
/m
B
T
viscosity parameter for temperature, T
B
vT
polymer viscosity parameter, T
1
c concentration (volume fraction)
C
1
ltration coefcient, L/L
C
se
effective salinity, N
C
st
static ltration parameter, L
3
t
0.5
D diameter, L
D
ch
polymer-chain diameter, L
D
hp
hydraulic polymer diameter (polymer size in solution), L
f
n
friction factor
f
X
fractional concentration of cation at the rock surface, N/N
h thickness, L
k permeability, L
k, k polymer-viscosity parameters, Eq. 15
k
sw1
clay-swelling parameters, Nt
1
k
sw2
clay-deswelling parameters, N
1
t
1
K
XC
cation-exchange parameter, N
z1
, where z is the valence
M
w
molecular weight, g/mol
n number
n
c
lter-cake compressibility, Lt
2
/m
n
F
dynamic-ltration exponent
N normality, meq/mL
p pressure, m/Lt
2
P

polymer shear-thinning exponent, Eq. 17


PI productivity index, rate/drawdown
PV pore volume, used as normalized volume unit
q ow rate, L
3
/t
RKA absolute permeability normalized on initial k
RKO oil permeability normalized on initial k
oi
RKOE k
o
at current maximum S
o
normalized on initial k
o
(maxi-
mum S
o
)
S skin factor (dimensionless), saturation
S
0
specic surface area (particle surface/particle volume), L
1
t time, t
T temperature, T
u Darcy velocity, u q/A, L/t
V volume, L
3
X ion concentration, N
Y erosion rate, L/t
z 1, ion valence
effective mud fraction in dynamic ltration
critical size parameter for particle ow in porous medium

nk
n 11, 22: trapping parameters for solids or polymer (k
s, p). n 41, 42: adsorption parameters for solids or
polymer (k s, p).

c
effective salinity parameter

e
erosion-rate parameter, t
1
shear rate, t
1

hf
polymer-viscosity parameter, t
1
[] intrinsic polymer viscosity, L
3
/m

nk
n 1, 2, 10, 20, 21, 22: trapping parameters for solids
or polymer (k s, p), L
1
. n 3, 4, 40, 41: adsorption
parameters for solids or polymers (k s, p), L
1
viscosity, m/Lt
dimensionless intrinsic viscosity
trapped material, PV, statistical variance
tortuosity parameter, (path length/length)
2

y
yield strength (lter cake), m/Lt
2
porosity
Subscripts
a average
c cake (lter cake)
cr critical
d dynamic
e end point
m mud
o oil
p particle, polymer
r relative
s solids
sp spurt loss
sw water in swelling clay
t trapping
w water
pore property
Acknowledgments
The major part of this work was conducted within the Well Pro-
ductivity 2002 Project supported by the European Union (EU)
Fifth Frame Programme. The authors would like to thank EU
and the industrial sponsors of the project, Agip BP Exploration,
ChevronTexaco, Norsk Hydro, and Shell Norway. We would also
like to thank the other partners in the project: ResLab, Eni Tecnol-
ogy, Rhodia, and M-I Drilling Fluids. Experimental data used in
this paper are supplied mainly by ResLab and Eni Tecnology. The
second phase of this work was supported by Shell, BP, and Statoil.
Supply of laboratory test data for model calibrations by Halliburton
is greatly appreciated.
References
Al-Abduwani, F.A.H., Hime, G., Alvarez, A., and Farajzadah, R. 2005.
New Experimental and Modeling Approach for the Quantification of
Internal Filtration. Paper SPE 94634 presented at the SPE European
Formation Damage Conference, Sheveningen, The Netherlands, 2527
May. doi: 10.2118/94634-MS.
Al-Abduwani, F.A.H., Shirazadi, A., van den Broek, W.M.G.T., and Currie,
P.K. 2003. Formation Damage vs. Solid Particles Deposition Profile
During Laboratory Simulated PWRI. Paper SPE 82235 presented at
the SPE European Formation Damage Conference, The Hague, 1314
May. doi: 10.2118/82235-MS.
Arthur, K.G. and Peden, J.M. 1988. The Evaluation of Drilling Fluid Filter
Cake Properties and Their Influence on Fluid Loss. Paper SPE 17617
presented at the International Meeting on Petroleum Engineering,
Tianjin, China, 14 November. doi: 10.2118/17617-MS.
Aziz, K. and Settari, A. 1979. Petroleum Reservoir Simulation. Essex, UK:
Elsevier Applied Science Publishers.
Bezemer, C. and Havenaar, I. 1966. Filtration Behavior of Circulating Drill-
ing Fluids. SPE J. 6 (4): 292298; Trans, AIME, 237. SPE-1263-PA.
doi: 10.2118/1263-PA.
Bourgoyne, A.T., Millheim, K.K., Chenevert, M.E., and Young, S.F. 1986.
Applied Drilling Engineering. Textbook Series, SPE, Richardson, Texas 2.
Cerasi, P., Ladva, H.K., Bradbury, A.J., and Soga, K. 2001. Measurements
of the Mechanical Properties of Filtercakes. Paper SPE 68948 presented
768 September 2010 SPE Journal
at the SPE European Formation Damage Conference, The Hague,
2122 May. doi: 10.2118/68948-MS.
Civan, F. 2000. Reservoir Formation DamageFundamentals, Modeling,
Assessment, and Mitigation. Houston, Texas: Gulf Publishing Company.
Dewan, J.T. and Chenevert, M.E. 2001. A Model for Filtration of Water-
Base Mud During Drilling: Determination of Mudcake Parameters.
Petrophysics 42 (3): 237250.
Ding, Y., Longeron, D., Renard, G., and Audibert, A. 2002. Modelling of
Both Near-Wellbore Damage and Natural Cleanup of Horizontal Wells
Drilled With a Water-Based Mud. Paper SPE 73733 presented at the
International Symposium and Exhibition on Formation Damage Con-
trol, Lafayette, Louisiana, USA, 2021 February. doi: 10.2118/73733-
MS.
Ferguson, C.K. and Klotz, J.A. 1954. Filtration from Mud During Drilling.
J Pet Technol 6 (2): 3043. SPE-289-G. doi: 10.2118/289-G.
Fordham, E.J., Allen, D.F., and Ladva, H.K.J. 1991. The Principle of a
Critical Invasion Rate and Its Implications for Log Interpretation. Paper
SPE 22539 presented at the SPE Annual Technical Conference and
Exhibition, Dallas, 69 October. doi: 10.2118/22539-MS.
Gruesbeck, C. and Collins, E. 1982. Entrainment and Deposition of Fine
Particles in Porous Media. SPE J. 22 (6): 847856. SPE-8430-PA. doi:
10.2118/8430-PA.
Hirasaki, G.J. 1982. Ion Exchange With Clays in the Presence of Surfactant.
SPE J. 22 (2): 181192. SPE-9279-PA. doi: 10.2118/9279-PA.
Hirasaki, G.J. and Lawson, J.B. 1986. An Electrostatic Approach to the
Association of Sodium and Calcium with Surfactant Micelles. SPE
Res Eng 1 (2): 119130; Trans, AIME, 281. SPE-10921-PA. doi:
10.2118/10921-PA.
Lake, L.W. 1989. Enhanced Oil Recovery. Englewood Cliffs, New Jersey:
Prentice Hall.
Meter, D.M. and Bird, R.B. 1964. Tube Flow of Non-Newtonian Polymer
Solutions: PART I. Laminar Flow and Rheological Models. AIChE
Journal 10 (6): 878881. doi: 10.1002/aic.690100619.
Outmans, H.D. 1963. Mechanics of Static and Dynamic Filtration In the
Borehole. SPE J. 3 (3): 236244; Trans., AIME, 228. SPE-491-PA.
doi: 10.2118/491-PA.
Pautz, J.F., Crocker, M.E., and Walton, C.G. 1989. Relating Water Quality
and Formation Permeability to Loss of Injectivity. Paper SPE 18888
presented at the SPE Production and Operations Symposium, Okla-
homa City, Oklahoma, USA, 1314 March. doi: 10.2118/18888-MS.
Pope, G.A. and Nelson, R.C. 1978. A Chemical Flooding Compositional
Simulator. SPE J. 18 (5): 339354. SPE-6725-PA. doi: 10.2118/6725-
PA.
Snchez, E., Audibert-Hayet, A., and Rousseau, L. 2004. Influence of Drill-
in Fluids Composition on Formation Damage. SPE J. 9 (4): 403410.
SPE-82274-PA. doi: 10.2118/82274-PA.
Semmelbeck, M.E., Dewan, J.T., and Holditch, S.A. 1995. Invasion-Based
Method for Estimating Permeability from Logs. Paper SPE 30581
presented at the SPE Annular Technical Conference and Exhibition,
Dallas, 2225 October. doi: 10.2118/30581-MS.
Sharma, M.M. and Yortsos, Y.C. 1986. Permeability Impairment Due to
Fines Migration in Sandstones. Paper SPE 14819 presented at the SPE
Formation Damage Control Symposium, Lafayette, Louisiana, USA,
2627 February. doi: 10.2118/14819-MS.
Sorbie, K.S. 1991. Polymer-Improved Oil Recovery. Glasgow, Scotland:
Blackie & Sons.
Suryanarayana, P.V.R., Wu, Z., and Ramalho, J. 2007. Dynamic Model-
ing of Invasion Damage and Impact on Production in Horizontal
Wells. SPE Res Eval & Eng 10 (4): 348358. SPE-95861-PA. doi:
10.2118/95861-PA.
van Oort, E., van Velzen, J.F.G., and Leerlooijer, K. 1993. Impairment by
Suspended Solids Invasion: Testing and Prediction. SPE Prod & Fac
8 (3): 178184; Trans., AIME, 295. SPE-23822-PA. doi: 10.2118/23822-
PA.
Wennberg K.E. and Sharma M.M. 1997. Determination of the Filtration
Coefficient and the Transition Time for Water Injection. Paper SPE
38181 presented at the SPE European Formation Damage Conference,
The Hague, 23 June. doi: 10.2118/38181-MS.
Wu, J., Torres-Verdin, C., Sepehrnoori, K., and Delshad, M. 2004. Numeri-
cal Simulation of Mud-Filtrate Invasion in Deviated Wells. SPE Res
Eval & Eng 7 (2): 143154. SPE-87919-PA. doi: 10.2118/87919-PA.
Appendix A
Computation of Average Filter-Cake Properties. The compressible-
cake permeability, k
c
(p
n
), can be expressed using Eqs. 5 and 10:
k p
e
e S
c n
o
n p
o
n p
c n
c n
( )

( )


3 3
2
0
2
2 1
. . . . . . . . . . . . . . . . . . . . . . . (A-1)
Substituting Eq. A-1 into Eq. 8 yields
k p
S
e
e
p
c n
p
o
n p
o
n p
n
c n
c n
d d
0 0
2
3 3
2
0
1
2
1

( )

. . . . . . . . . . . . . . . . . . (A-2)
Defining z e
n p
c n


1 1 in Eq. A-2 yields
k p
S n
z
z
z
k p
c n
p
c z
z
c n
p
d d
d
0 0
2
2
2
0
1
2
1
1
2

( )


+
j
(
,
\
,
(

1
2
1
2
0
2
0
1
2
S n z
z z
c z p
z p p
ln
( )
( )
. . . . . . . . . . . . . . . (A-3)
Substituting the left-hand side with k
ca
p (Eq. 8) and using
F p
z
z z
k n
( ) +
1
2ln , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-4)
the average filter-cake permeability is expressed by
k
S n
F F p
p
ca
c
k k

( ) ( ) 1
2
0
0
2

. . . . . . . . . . . . . . . . . . . . . . (A-5)
Let k
ca
(x) k
ca
[p
n
(x)] be the average permeability from the cake
surface to depth x within the cake. Because the rate and viscosity
of the permeate flow are constant, the product k
ca
[p
n
(x)]p
n
(x)/x
must be constant. If we define x as the relative position within the
cake as a function of p
n
, then the distance from the cake surface
to pressure p
n
(x) is
x p
h
h
F p F
F p F
n
c
c
k n k
k k
( )
( ) ( )
( ) ( )

0
0
. . . . . . . . . . . . . . . . . . . . . (A-6)
In order to calculate the flow rate, we also need to know the
thickness (d
c
) of the cake. This is readily obtained from mass bal-
ance once the average porosity of the cake is found. Using Eqs. 7
and 8 we can write the average porosity as

c
c n
p
c n
p
x
k p
k p
d
d
d
0
0
1
0

. . . . . . . . . . . . . . . . . . . . . . . . . . (A-7)
The denominator is given in Eq. A-3, and the numerator integral
is after using the previous substitution, z 1:

c n
p
c z
k p
n S z
z z z d
0 0
2
1
2
2
0
1
2
1
3 3



+ +
j
(
,
\
,
(
ln
(( )
( ) z p
. . . . . . . . . . (A-8)
By using
F p
z
z z z
n
( ) + +
1
3 3
1
2
2
ln , . . . . . . . . . . . . . . . . . . . . . . (A-9)
we obtain the average porosity as

( ) ( )
( ) ( )
F p F
F p F
k k

0
0
. . . . . . . . . . . . . . . . . . . . . . . . . . . (A-10)
September 2010 SPE Journal 769
Arild Lohne is a senior research engineer in the improved oil
recovery (IOR) group at the International Research Institute of
Stavanger (IRIS). His research interests include water-based IOR
methods and multiphase upscaling. He holds a BS degree in
environmental engineering and an MS degree in petroleum
technology, both from the University of Stavanger. Liqun Han
is a senior research scientist with Halliburton Energy Services,
Stavanger, specializing in reservoir drilling fluids and well pro-
ductivity. Before join Halliburton in 2006, he worked for Rogaland
Research Institute (now IRIS) for almost 11 years. He holds a
BS degree from the University of Petroleum, China, and a PhD
degree from Heriot-Watt University, Scotland, both in petroleum
engineering. Claas van der Zwaag is a leading advisor in res-
ervoir drilling and completion fluids at Statoil and has a part-
time assignment as adjunct at the Deptartment of Petroleum
at the University of Stavanger. Van der Zwaag holds a Diploma
Engineering degree in mining engineering from the RWTH Aachen
University, Germany, and a Doctor of Engineering degree in drill-
ing engineering from the University of Science and Technology
in Trondheim, Norway. Hans van Velzen is production chemist for
NAM/Shell. During 30 years with NAM/Shell, he has been involved
in many research projects. Currently, he is well fluids engineer at
NAM in Assen. Van Velzen holds an HTS degree from Haagse
Hogeschool, The Hague, in chemical technology. Anne-Mette
Mathisen is a leading advisor in production technology at
Statoil. She holds an MS degree in petroleum technology from
the University of Stavanger. Her focus areas are well productivity,
formation damage, sand control, and perforation technology.
Allan Twynam is a senior drilling engineer at BP. Wim Hendriks
is a senior production technologist at Shell in Assen. Roman
Bulgachev is currently drilling fluids and formation damage
engineer and laboratory projects supervisor at BP Exploration,
UK. He joined BP in February 2007. Before that, Bulgachev was
Halliburton-Baroid Drilling Fluids project manager and drilling flu-
ids field engineer in Russia for 5 years. He holds an MS degree
in chemistry from Lomonosov Moscow State University. Dimitrios
G. Hatzignatiou is chief technical director at IRIS and adjunct
professor at the University of Stavanger. Before that, he was a
reservoir engineering advisor, held several management and
principal technical positions with Schlumberger, and was also
a professor of petroleum engineering at the University of Alaska.
He holds a PhD degree in petroleum engineering from the
University of Tulsa, and his areas of specialization are reservoir
engineering, IOR, reservoir characterization, production opti-
mization, reservoir simulation, reservoir management, and CO
2

sequestration. Hatzignatiou is a registered professional engineer
in Europe and serves as an associate editor of the SPE Reservoir
Evaluation and Engineering journal.

Вам также может понравиться