Você está na página 1de 20

SPE 122241

Formation Damage and Well Productivity Simulation


Arild Lohne, SPE, IRIS; Liqun Han,* SPE, and Claas van der Zwaag, SPE, StatoilHydro; Hans van Velzen,
Shell/NAM; Anne-Mette Mathisen, SPE, StatoilHydro; Allan Twynam, SPE, BP; Wim Hendriks, SPE, Shell;
Roman Bulgachev, SPE, BP; and Dimitrios G. Hatzignatiou, SPE, IRIS
*Formerly at IRIS, now with Halliburton

Copyright 2009, Society of Petroleum Engineers

This paper was prepared for presentation at the 2009 SPE European Formation Damage Conference held in Scheveningen, The Netherlands, 27–29 May 2009.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract and more realistic pre-screening of drilling fluids than


In this paper we describe a simulation model for computing traditional core plug testing.
the formation damage imposed on the formation during over-
balanced drilling. The main parts modelled are filter cake Introduction
build-up under both static and dynamic conditions, fluid loss A simulation tool, referred to as Maximize, has been
to the formation, transport of solids and polymers inside the developed for the purpose of investigating fluid loss to the
formation including effects of pore lining retention and pore formation during over-balanced drilling and the impairment
throat plugging, and salinity effects on fines stability and clay imposed on the formation by the invading fluid. The
swelling. The developed model can handle multi-component objective of the program is to serve as a tool for supporting
water-based mud systems at both the core scale (linear well planners’ decisions related to the choice of well fluids,
model) and the field scale (2D radial model). Among the and integrating, analyzing and interpreting laboratory as well
computed results are fluid-loss versus time, internal damage as field formation damage data.
distribution and productivity calculations for both the entire The filtration properties of the mudcake forming at the
well and individual sections. wellbore surface has been investigated by several authors
The simulation model works in part independent on fluid over the years, see for example Ferguson and Klotz (1954);
loss experiments, e.g., we do not use fluid leakoff Outmans (1963), Bezemer and Havenaar (1966), Arthur and
coefficients, but instead we compute the filter cake buildup Peden (1988), Fordham et al. 1991 and Dewan and
and its flow resistance from properties ascribed to the Chenevert (2001). The common understanding is that once
individual components in the mud. Some of these properties the filter cake has been formed, it will control the filtration
can be measured directly, such as particle size distribution of rate independent of the formation properties, except at very
solids, effect of polymers on fluid viscosity and formation low permeability where the flow resistance offered by the
permeability and porosity. Other properties, which must be formation is comparable to the filter cake resistance.
determined by tuning the results of the numerical model The filter cake properties depend only on its composition,
against fluid loss experiments, are still assumed to be rather the pressure drop over the cake Δp and the shear stress acting
case independent, and once determined they can be used in on the cake surface by the circulating mud. Under dynamic
simulations at altered conditions as well as with different conditions, the filtration rate will approach asymptotically a
mud formulations. A detailed description of the filter cake limiting steady-state rate which only depends on the shear
model is given in the paper. stress at the cake's surface.
We present simulations of several static and dynamic A combined filter cake and simulation model was used by
fluid loss experiments. The particle transport model is used to Semmelbeck et al. (1995) for computing the fluid invasion
simulate a dilute particle injection experiment taken from the profile along the well. Further improvement of the filter cake
literature. Finally, we demonstrate the model’s applicability model was presented by Dewan and Chenevert (2001), who
at the field scale and present computational results from an demonstrated the derived model's capability to reproduce
actual well drilled in the North Sea. These results are complex laboratory experiments with sequential changes in
analysed and it is concluded that the potential impact of the dynamic shear rate and overbalance pressure. Others, have
mechanistic modelling approach used is (a) increased also investigated filtrate invasion by numerical simulations
understanding of damage mechanisms, (b) improved design e.g., Ding et al. (2002), Wu et al. (2004) and Suryanarayana
of experiments used in the selection process and (c) better et al. (2005).
predictions at the well scale. This allows for a more efficient
2 SPE 122241

Knowledge of the filtrate profile along the wellbore is models (appropriate for the field scale). We use two
important for well log interpretations and for evaluating the boundaries (wells) constrained by constant rate or constant
damage imposed on the formation by the invading fluid. A pressure. The flow equation is solved using the IMPES
number of potential damaging mechanisms and modelling of method (implicit in pressure and explicit in saturations with
these are described in details by Civan (2000). For water- respect to time). The reader is referred to the literature, e.g.,
based muds, one of the main damage mechanism is Aziz and Settari (1979) for further details on reservoir
particulate plugging of the formation, either by externally simulators.
introduced particles or by in situ fines mobilized by the The phase flow between grid cells is computed from the
invading fluid. The mechanisms of particle transport in pressure solution. Knowing the flow, local concentrations of
porous medium have been investigated by e.g., Gruesbeck oil, water and dissolved or suspended species are updated for
and Collins (1982), Sharma and Yortsos (1986), Wennberg the new time-step. Chemical and physical processes are
and Sharma (1997) and Al-Abduwani et al. (2005). decoupled from the flow equation. Components are first
Traditionally, the computation of fluid loss has relied on moved between cells, and then changes over the time-step,
using experimental filtration rates or, as in the more due to physical and chemical process, are computed locally
sophisticated models, experimental cake permeability and in each grid-cell.
porosity obtained for a specific mud. In our approach, we Rock properties such as absolute permeability k, porosity
bring this one step forward by computing the cake properties φ and also relative permeabilities and clay content are
from its compositions. The properties from which the cake's allowed to vary between layers (sections). Moderate
porosity and permeability are computed are ascribed to the compressibility of both the formation and the fluid
individual components. In Ding et al. (2002) and components is handled by the model. Corey type relative
Suryanarayana et al. (2005) the damage in invaded zones are permeabilities are used:
simulated based on experimental return permeability and an
assumed damage profile. We make no such assumptions, but S j − S jr
k rj = k rje S Ej
jn , S jn = , j = o, w . (1)
model the flow and retention of both particles and polymers 1 − S wr − S or
inside the formation, and compute the permeability reduction
from the amount of trapped material. As stated earlier, this is a multi-component simulator.
Models that are used should be consistent with the Properties of components, like viscosity or their effect on
physics involved at a macroscopic level. Input parameters viscosity, are described in the following sections.
should ideally be of three kinds, (a); properties of involved In simulations with mud filtration, the rate will change
components (e.g., viscosity, density, size of particles,..), (b); very rapidly during the first period when the filter cake is
properties of the formation (e.g., permeability, porosity, ....) built up from zero thickness, and an IMPES-type solution
and (c); model parameters describing events which are will be unstable. We solve this by pre-computing the growth
independent of the above properties (a and b). The benefit of of the cake during the time-step assuming the pressure
this approach is that the amount of empirical input difference between the well (mud pressure) and the
parameters are reduced and that the parameters used are more connection block to be constant. The estimated cake
universal and show less variance among different properties are used in the computation of the new pressure
experiments. Finally, the proposed approach will increase the solution for the total grid, and finally the rates from this
utilization of earlier experience since ideally all experiments solution are used to update the real growth of the filter cake.
should be matched by one single data set, provided adequate
descriptions of the experiments are available. Filter cake
The filter cake model presented here computes cake
Simulation model properties from properties of the individual components
This section describes the various modules that constitute the present in the mud. The model handles an arbitrary number
newly developed simulation model. The two main parts of of polymers and solids.
the Maximize program are the filter cake model handling Static filtration. The relation between cake thickness hc
filter cake buildup and controlling flow into the formation and filtrate volume Vf is obtained from mass balance
and the reservoir flow model handling flow inside the Bourgoyne et al. 1986:
formation. The fluids introduced to the formations contain a
V f C1 c sm
number of dissolved and dispersed components, which in hc = , C1 = , (2)
turn may change the original flow properties of the formation Ac c sc − c sm
through different chemical and physical processes. The
where Ac is the filter cake area and csm and csc are the
retention of solids and polymer are split into a pore throat
concentrations of solids in the mud and the filter cake,
trapping model and pore lining adsorption model. Brine
respectively. The flow rate through the cake expressed by
interaction with the rock surface and clays are described by a
Darcy's law is:
multi-component cation exchange model, which provides
(3)
input to processes like fines migration and clay swelling. ΔpAc k c
q= .
μhc
Flow model
The flow model handles two-phase flow in 1-D linear Setting the rate q = dVf/dt, inserting hc from Eq. 2 and
(typically used for cores), 2-D rectangular and 2-D radial integrating the resulting expression yields the known filtrate
SPE 122241 3

volume versus square root of time relation (a spurt loss The local pressure gradient within the filter cake can be
volume Vsp is also added to the resulting expression): described by Darcy's equation. At a given time, the flow rate
u will be constant through all layers of the cake for an
2Δpk c incompressible fluid. Then, according to the Darcy equation,
V f = Vsp + Ac C st t , C st = . (4)
μC1 the term kcdp/dx must be constant (independent on location in
the cake) and equal to u·μ. (assuming μ is constant):
There are three unknowns in Eq. 4, namely the spurt loss Vsp,
the cake permeability kc and the cake porosity φc (csc =1-φc is dp n dp uμ
=− = . (7)
part of C1). Two of these variables, kc and φc, depend on the dx dx k c ( x)
mud composition and the pressure drop over the cake while
the spurt loss depends on the mud composition and properties Let pn represent the local net overburden pressure in the
of the rock. The cake's porosity and permeability will cake (pn=pm-p) and Δp the fluid pressure drop over the cake.
decrease with increasing pressure drop, due to The overburden pressure pn will range from zero at the cake
compressibility of the cake, and thus reduce the rate increase. inlet to Δp at the cake outlet. The velocity u can be expressed
The main effect of temperature is through its influence on in terms of the average cake permeability kca and Δp. Let kc
the filtrate viscosity; and thereby on the filtration parameter be a function of pn and combine kca with Eq. 7 gives:
Cst. Different polymers used in mud will have different upper Δp hc
k Δp
temperature limits above which they become chemically
∫ k c ( p n )dp n = uμ ∫ dx, u = ca ,
unstable. Below this limit, only a weak temperature
0 0
μhc
dependency is assumed.
Δp (8)
∫ k c ( p n )dp n
Cake permeability. The permeability of a porous medium
can be estimated with the Carman-Kozeny equation (Lake
1989). Because mud is a multi-component system containing k ca = 0 .
different kinds of solids of very different sizes, we choose the Δp
following variant: The effective permeability will be only a function of Δp
∑ S 0i ci
and independent of the cake thickness. To solve the integral
φ3
k= , S0 = , (5) of kc(pn) we need a functional relation between kc and pn. In
2τ (1 − φ ) S 0
2 2
∑ ci the literature, kc and φc are normally treated uncoupled using
the same equation type (Outmans 1963):
The tortuosity parameter τ = (Lt/L)2 accounts for the
effective flow path length Lt, ci is volumetric concentration of k c = k o p − n1 and φ c = φ o p − n2 , n1 , n 2 ≥ 0 . (9)
species i and S0 is the specific surface area of the medium
In our case, we want to calculate the permeability using
defined as the surface to volume ratio. In particular for
the Carman-Kozeny Eq. 5, include the effect of
spheres, S0=6/Dp, where Dp is the particle diameter. The
compressibility through the porosity and integrate kcdpn to
specific surface area in a mixed system is readily obtained by
obtain the average kca. We use the classic definition for
volume weighted averaging. The operator <·> used for S0
compressibility, i.e.,
indicates that the property represents a volume weighted
average. 1 dφ
Another useful expression that relates an effective pore nc = − ⇒ φ = φo e −nc pn , nc ≥ 0 . ( 10 )
φ dp
diameter to permeability and porosity is the following:
Note that the porosity in Eq. 10 converges to the correct
32τk limit φο when pn approaches zero, which is not the case with
Dφ = . (6)
φ Eq. 9. The solution to the average cake permeability, as well
as the computation of the average porosity, is given in the
The Carman-Kozeny approach treats the porous medium Appendix. Here we show only the final expressions. Defining
as a bundle of capillary tubes and Dφ corresponds to the tube z(pn) = 1-φ(pn), the average permeability of a compressible
diameter when all tubes have the same diameter. filter cake is:
Cake compressibility. The compression of the cake is a
result of the local pressure difference between the cake 1 Fk (Δp ) − Fk (0)
k ca = ⋅ ,
matrix and the fluid filling the pore space. The cake matrix 2τS 02 nc Δp
pressure is assumed constant and equal to the mud pressure ( 11 )
1
pm. The fluid pressure decreases through the cake as a result Fk ( p n ) = + 2 ln z − z.
of the viscous flow. Since this pressure difference will vary z
throughout the cake, physical cake properties such as
porosity and permeability, which depend on the “overburden and the average filter cake porosity is:
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.
4 SPE 122241

Fφ (Δp ) − Fφ (0)
Polymer viscosity. The effect of polymer concentration on
φ = , the fluid viscosity is typically expressed at low shear rate
Fk (Δp ) − Fk (0) with a modified Huggins equation Sorbie (1991). We use the
( 12 ) expression:
1
Fφ ( pn ) = + 3 ln z − 3z + 12 z 2 .
z (
μ 0 = μ s 1 + νc + k 'ν 2 c 2 + k ' 'ν 3c 3 , ) ( 15 )
The average filter cake permeability and porosity where c is the polymer concentration (volume fraction), μs is
computed by Eqs. 11 and 12 can be used directly in Eqs. 2 the solvent (e.g., water) viscosity and ν is a dimensionless
and 4. form of the intrinsic viscosity [η] in units [cm3/cm3]. The
The compressibility of the filter cake can be split into a
intrinsic viscosity [η] is defined as the limit of the reduced
reversible (elastic) and an irreversible (compaction) part. A
viscosity when polymer concentration approaches zero. k’ is
simple way to model this is to apply an irreversibility factor, the Huggins constant, which for a range of polymers in good
Firf. Let Δpmax be the historical maximum Δp. If the current
solvents is reported to be equal to 0.4±0.1. Eventual changes
pressure drop Δp is less than Δpmax, then an apparent effective in temperature and in molecular weight Mw due to, e.g., pore
pressure drop Δpa is used: throat trapping is accounted for by the following expression:
Δpa = Δp + Firf(Δpmax - Δpc). ( 13 )
⎛ M ( x, t ) ⎞
a
Partially reversible compression of a filter cake is ν = [η ]ρ p ⎜⎜ w ⎟⎟ (1 − BvT (T − T0 )) , ( 16 )
illustrated in Figure 1. ⎝ M w0 ⎠
Solids. The information needed in the filter cake model
includes a size distribution of the solid particles, the specific where exponent a is reported to be in the range 0.5-1 (higher
surface area S0 and the solids contribution to the cake's value for good viscosifiers). T and T0 are actual and reference
porosity and compressibility. temperatures, respectively, and BνT is an empirical parameter.
The size distribution can be either in the form of a The units for the density ρp must be consistent with the units
lognormal distribution or entered as a table. If a tabular form used for [η] so that ν becomes dimensionless. The shear
is used, the volumetric cumulative distribution function thinning effect of polymers is represented by the Meter's
F=[0,1] is taken to be piecewise linear when plotted against model (Meter and Bird 1964) with the final polymeric
ln(D). S0 is computed by integrating S0(D) over the size viscosity given by:
distribution. If tabular input with n rows is used: μ0 − μ s
μ p = μs + ,
6φ Pα −1
S 0 (D ) = sh , ⎛ γ
1+ ⎜

⎟ ( 17 )
D ⎜ γhf ⎟
⎝ ⎠
xi
6φ sh ⎛ 1 1 ⎞
ΔS i = ∫ x
ci dx = 6φ sh ci ⎜⎜ − ⎟⎟, where exponent Pα has a value between 1 and 2 (high shear
xi −1 e ⎝ Di −1 Di ⎠ thinning) and γhf is the shear rate γ at which (μ0-μs) is
( 14 )
F − Fi −1
ci = i , x = ln (D ), reduced to one half. The main influence of temperature is
xi − xi −1 through the solvent viscosity, μs. We use the same Arrhenius
n type equation for oil and water to compute the viscosity as a
S 0 = ∑ ΔS i . function of pressure and temperature:
2 ⎡ ⎛1 1 ⎞ ⎤
⎢ BT ⎜⎜ − ⎟+ B p ( p − po )⎥
⎟ ( 18 )
The shape factor φsh is defined as the ratio of the particle's μ = μoe ⎣⎢ ⎝ T To ⎠ ⎦⎥ ,
surface area to the surface area of a sphere with the same
volume, i.e. for spheres φsh=1. Examples for other shapes; φsh where μo is the reference viscosity at temperature To and
is 1.24 for a cube and 2.3 for a flat particle with dimensions pressure po. BT and Bp are empirical constants. The
0.1×1×1 in any units. The density probability function f(D) = temperature is in absolute units (°K). We use the fluid
dF/dD. viscosity in Eq. 18 to describe flow through the filter cake,
For each solid type present in the mud, a reference Eqs. 3 and 4. Yet another correction for the content of solids
porosity φo corresponding to zero overburden pressure and a is done before the effect of polymer is computed:
compressibility parameter nc are required.
μo
Polymer. For polymers we need to know their effect on μs = . ( 19 )
fluid viscosity, which is important in dynamic filtration and (1 − csm )2.5
for flow inside the porous medium. Secondly, we need the
size of the polymer molecules in order to evaluate Polymer size. We assume a lognormal distribution of Mw
entrapment in the filter cake and in the formation. Finally, we in the computations. The size of the polymer molecule in
must estimate the resistance to flow caused by the polymer in solution can be estimated from the intrinsic viscosity [η]
the filter cake. using Flory's empirical equation for the mean end-to-end
distance:
SPE 122241 5

Dhp = 0.00017(M w [η ])1 3 . ( 20 ) hsp c sm*


Vsp = Ac , C1sp = , ( 24 )
C1sp c sc* − c sm*
The molecular weight Mw is in g/mole, [η] in ml/g and the
computed polymer diameter in µm. We will use the hydraulic where csm* is the concentration of solids with D>Dcrs in the
polymer diameter Dhp for computing polymer entrapment in mud and csc* is equal to (1-φo) computed for solids alone. In
the filter cake and pore throat trapping of polymer inside the the special case of spherical and monosized particles, nsp=1
formation. will correspond exactly to a single monolayer with porosity
Flow resistance. The use of the Carman-Kozeny Eq. 5
φo. A value in the range 1-2 for nsp is suggested.
requires the polymer's contribution to the specific surface
Filtration period. In this period, all solids are retained in
area and the cake porosity. The surface to the polymer is
the cake. Deposited material during the spurt loss period is
estimated by assuming the molecule to be a long rod with
ignored, i.e., cake thickness starts from zero. The first thing
diameter Dch. The specific surface area is then:
to evaluate is the retention of polymer in the cake. The part
4 of the polymer molecules with Dhp smaller than a critical size
S0 p = . ( 21 ) Dcrp will pass through the cake. Dcrp is estimated as a fraction
Dch
αp of the pore size Dφc (Eq. 6) for a filter cake made up of
The specific surface area for a given polymer type will solids alone.
depend on the chain diameter Dch of the repetitive unit, but
Dcrp = α p Dφc . ( 25 )
not on the molecular weight.
For each polymer type i we specify a φoi and nci according The effective polymer concentration that will be retained
to Eq. 10 that will go into the computations of the effective in the cake is then obtained from the Mw size distribution.
parameters for the cake. The critical size ratio αp represents the average probability
Cake build-up. The cake build-up calculations involve for entrapment over a long filtration period and should be in
two periods, the spurt loss period and the filtration period the low range of standard trapping rules (1/7-1/3). A value of
after a filter cake has been formed. 0.1 has been used in the simulations presented here.
Spurt loss period. In the spurt loss period, all polymers Finally, the cake properties can be computed. First,
and solids particles smaller than a critical size Dcrs are effective cake porosity, compressibility and specific surface
assumed to enter the formation. Solid particles larger than area are estimated as the volume weighted average of all
Dcrs will be held back on the rock surface. When sufficient involved components, solids and polymers (nt components).
coverage of the rock surface is reached, all solids will be
retained. The critical particle size for solids is computed nt nt
from: ϕo c
= ∑ c c,iϕ o,i , nc c = ∑ c c,i nc,i ,
i =1 i =1
Dcrs = α s Dφ , ( 22 ) nt nt
( 26 )
S 0 c = ∑ cc,i S 0,i , ∑ cc,i = 1.
where Dφ is obtained by Eq. 6 and parameter αs represents i =1 i =1
the lower size ratio between particle size and pore size that
will result in trapping at the rock surface. Note that Dφ is an where cc,i represents the relative concentration of component
average property representing both pore bodies and pore i in the filter cake matrix. The average cake permeability and
throat distributions. Some particles larger than Dcrs will enter porosity are readily obtained from Eqs.11 and 12, and finally
the rock and some smaller particles will be held back. What the static filtration parameter Cst in Eq. 4 can be computed.
we want is to compute the average fraction of particles held Dynamic filtration.The filter cake model outlined so far
back during the short spurt-loss period. From our experience applies for static conditions at the filter cake surface. In real
situations the filter cake surface is exposed to shear forces
so far, a value for αs in the range of 0.3-0.5 seems to be
from mud circulating through the well. Two main
appropriate.
mechanisms are considered in the dynamic filtration model,
The average height of the layer needed to form a covering
namely reduced attachment and erosion.
layer on the rock surface is computed with:
The flow along the cake surface will cause part of the
6 particles to roll off the surface and bounce back into the mud.
hsp = n sp hcrs , hcrs = ,
S 0 (D > Dcrs )
( 23 ) Only a fraction α of the particles will stick to the cake
surface. This will reduce the growth rate of the cake. The
where hcrs represents a characteristic height of the retained shear stress at cake surface may also detach particles from
particles computed from the specific surface area for the the filter cake and thus decrease the cake thickness. This is
fraction having D>Dcrs. nsp is the number of "layers" needed termed the erosion rate Y. The cake growth in terms of α and
to form a completely covering cake. Note that this way of Y can be obtained from mass balance:
computing hsp puts more weight on the smaller particles and
reduces the spurt loss for flattened particles. The spurt loss
volume is then:
6 SPE 122241

dhc ⎛ Y ⎞ σ z = γμ z , μ z = μ (γ ) . ( 33 )


= C d1 ⎜⎜ u x − ⎟,

dt ⎝ α c sm ⎠ Assume a multi-layered cake, with layer thickness Dp (a
( 27 )
c smα characteristic particle diameter size). The forces holding the
C d1 = . layer in place is cohesive attraction between the particles
c sc − c smα
(solids and polymers) that form the cake matrix and the
The flow through the cake ux is given by Darcy's law if viscous pressure gradient from the flow through the cake. We
the cake permeability and current thickness are known. This assume cohesive forces to be represented by the filter cake
leaves two unknowns to be determined, α and Y. yield strength, τy, which is measured by Cerasi et al. (2001)
Consider a non-moving spherical particle just in contact to range from a few 100s Pa to a few 1000s Pa. The pressure
with the filter cake surface with a filter loss rate ux and a drop over the first layer will be equal to the net overburden
perpendicular mud flow along the surface, uz. For simplicity, pressure at distance Dp into the cake, pn(x=Dp), which we
we assume a constant shear rate γ near the surface, so that obtain from Eq. A.6. The x-directed forces considered are
the average velocity uz in a layer, with the same thickness as then:
the particle, will be: ( )
σ x = pn x = D p + τ y . ( 34 )
du z
= γ, u z ( x) = γx We assume the erosion rate to be proportional to the ratio
dx of forces acting in z- and x-directions, and to Dp. An
Dp . ( 28 )
1 empirical rate constant βe (in units s-1) is introduced. The
uz =
Dp ∫ γxdx = 12 γD p erosion rate is given by:
0
γμ z
Y = βeD p
p n (D p ) + τ y
The force acting on a spherical particle with constant . ( 35 )
relative velocity in a fluid is given by:
F = 3πμuD p . ( 29 ) The particle diameter used in Eqs. 28 to 35 is estimated
from the size distribution of both solids and polymer using
Then for the particle close to the surface, the forces in x volume weighted D-1.
and z directions are: A limitation of the filter cake model is that we are using
average cake properties. This prevents us from properly
Fz = 32 πμ (γ )γD 2p including physical processes like size selective attachment to
. ( 30 ) the cake and pressure dependent yield strength, τy. A
Fx = 3πμ o u x D p reasonable implementation of such effects requires some
kind of spatial discretisation of the filter cake.
Note that in Fz, the viscosity μ will be a function of the
shear rate, while the viscosity acting towards the surface (in
Particle retention in the formation
Fx) is computed at zero shear rate (μο). The dominating mud
Retention of particles flowing through a porous medium
flow will be along the surface. The reduced viscosity at
can be divided into two main parts, pore throat plugging and
higher rates (shear thinning) is due to polymer molecules
pore line filling. The retention kinetics for the two parts is
orientating themselves along the main flow direction.
very different and so is the damaging effect. In the case of
Particles bouncing back into the mud will not see this shear
external particles (injection of suspended particles), pore
thinning, as they will be moving perpendicular to the flow.
throat trapping will be most important close to the inlet,
The ratio of forces acting on the particle along and towards
while pore lining will dominate deeper in the core. Larger
the surface is then:
particles will be filtered out from the solution as they are
Fz μ (γ )γD p trapped in narrow pore throats, while smaller particles can
FR = = . ( 31 ) penetrate deeper. In laboratory experiments, normally only
Fx 2μ o u x the combined effects of the two processes are obtained,
The fraction of solids (in the mud) that will attach to the resulting in a large variety of retention models/parameters.
The observed results will strongly depend on the
surface (α) is modelled by:
experimental setup and there is no good method for scaling
1 Fz μ (γ )γD p the experimental results to other conditions.
α= , FR = = . ( 32 ) Here, we will try to separate the two retention types. Our
1 + FR F
n f n Fx f n 2μ o u x
main focus is on the pore throat trapping, as it has the higher
The exponent nF determines how fast the transition of a is damage potential.
when the filtration rate ux decreases and its value should Pore throat trapping. In pore throat trapping, only
reflect the width of the particle distribution. Values in the particles large enough to be held back in the narrow parts of
range 0.5-1 are used in the simulations here. Finally, fn is a the pore space (throats) will be involved. To model this, we
friction factor. need to keep track of particle sizes. Solid and polymer
The shear stress acting on the filter cake surface is: components are split into a number of size fractions, each
represented as a monosized sub-component in the simulation.
SPE 122241 7

The basic equation used for both solids and polymers (k=s,p) involved. The likely explanation is that small amounts of
allows for both trapping and re-entrance of particles back to trapped material will not be packed close together as a cake,
solution. The trapping rate for sub-component i is computed but rather be spread over a larger area within the narrow parts
by: of the pore space. One way to compensate for this is to
introduce a modification factor for the specific surface area
dσ i
= w (λ1,k ct ,i − λ2,k σ i ) , k=s,p.
u of the polymer, fpm.
( 36 )
dt φS w In the second method we compute a "pore throat
permeability" kpt within the narrow pore space defined as a
The units for trapped material and total mobile fraction fpt of the total pore space. A specific surface S0pt for
concentration of component i is in [pv] units so that over a this pore throat volume is computed by combining surface
time step Δt we have Δσi=-Δct,i. The unit for parameters λ1 area of the original rock, S0r(φ, k) obtained from Eq. 5, with
and λ2 is 1/length. For solids, the de-trapping term λ2 will be the surface area of the deposited solids and polymers. The
(close to) zero. If λ2 is zero, Eq. 36 says that a fraction of all rock porosity is corrected for deposited material, φpt. Then
passing particles equal to λ1 will be retained per length unit kpt(φpt, S0pt) is calculated by Eq. 5 and finally the new
and the expected traveling length E[x] is 1/λ1. To make the effective permeability is obtained by harmonic averaging
equation more general, we allow λ1 to be a function of with k.
particle size and λ2 to depend on fractional flow. For solids,
λ1 and λ2 are given by: ⎛ σ p +σs ⎞
φ pt = φ ⎜⎜1 − ⎟,

α11s ⎝ a ⎠
⎛ Di ⎞
λ1s = λ10 s ⎜
⎜ Dφ


, λ2 s = λ20 s (1 + λ21s f o ) , ( 37 )
S 0 pt =
( )
S 0 r (1 − φ )a + S 0 sσ s + S 0 pσ p φ
(1 − φ )a + (σ s + σ p )φ
⎝ ⎠ ,

if the sub-component size Di > αtsDφ. If the particle size is −1 ( 40 )


⎛ a 1 − a ⎞⎟
k eff = ⎜
smaller than this limit, trapping is not considered. Finally, Dφ
+
is the pore size estimated by Eq. 6.
Contrary to rigid solid particles, polymer particles will be ⎝ (
⎜ k pt φ pt , S 0 pt ) k ⎟⎠
,

quite deformable, and their retention in pore throat may be ⎛ σs +σ p ⎞


a = max⎜⎜ f pt , ⎟.
1 − φic ⎟⎠
highly dependent on the viscous drag from fluid passing
through. For polymer, λ1 and λ2 are computed by: ⎝
α11 p Pore lining retention. The pore lining retention concerns
⎛ d hp,i ⎞
λ1 p = λ10 p ⎜ ⎟ , material attached to the pore surfaces. The attachment of
⎜ Dφ ⎟ particles to the rock surface is governed by a balance
⎝ ⎠
( 38 ) between attractive and repulsive forces. This is modelled
⎛ ⎛ α 22 p ⎞
⎜ ⎞ ⎟
( )
∇p with a Langmuir type kinetic equation for adsorption (A).
λ2 p = λ20 p 1 + λ21 p f o ⎜1 + ⎜ ⎟
⎟.
⎜ λ ⎟
= λ3, k ck (Qm, k − Ak ) − λ4, k Ak , k=s,p.
⎜ ⎝ 22 p ⎠ ⎟ dAk
⎝ ⎠ ( 41 )
dt
The size of the polymer sub-component i is represented
by the hydraulic diameter from Eq. 20 and a lower size limit The same equation is used for both polymers and solids.
for pore throat trapping can be specified by αtpDφ. Particles Qm is the adsorption capacity, c the volumetric concentration
which are smaller than this limit will not be trapped in the and λ3 and λ4 are rate parameters representing adsorption and
simulations. The main differences from solids is that the re- desorption respectively. For a mixture with several solid and
entrance term can not be ignored, and that the release rate polymer components, the following apply:
will increase with increasing pressure gradient, ∇p . • One set of polymer adsorption parameters is used for all
polymers. The maximum adsorption capacity for
Two methods for computing the permeability reduction
individual polymer components is according to its
caused by the pore throat trapping are implemented. In the
relative concentration and sums up to Qm,p. Adsorption
first, an internal cake permeability kic is computed similarly
between polymers are competitive.
to kc for the filter cake, except that a zero overburden
pressure pn is used. Then the effective absolute permeability • One set of solid adsorption parameters is used for all
is computed by harmonic averaging between kic and the solids. The maximum adsorption capacity for individual
original rock permeability k. solid components is according to its relative
concentration and sums up to Qm,s. Adsorption between
−1
⎛ a 1− a ⎞ σ +σ p solids are competitive.
k eff = ⎜⎜ + ⎟⎟ , a = s . ( 39 ) • The model assumes that polymer and solid adsorption
⎝ k ci k ⎠ 1 − φic can be decoupled, i.e., that no competitive adsorption
between solid and polymer exists.
The internal cake porosity is weighted average <φo> of all A main reason for repulsive forces is the electrical charge on
trapped material. This method seemed to over-predict the rock and particle surfaces. An increasing electrolyte
permeability reduction, in particular when polymer is concentration will reduce electrostatic repulsion. Also, a
8 SPE 122241

particle may departure from the rock surface as a result of a


dσ sw k sw1
high viscous drag. These mechanisms are included in the = (Qsw max − σ sw ) − k sw2C seσ sw , ( 45 )
desorption parameter λ4 through the pressure gradient ∇p dt C se + ε cse
and effective salinity Cse terms. where σsw is the swelling state defined as the volume ratio of
α 41,k adsorbed water to clay, ksw1 and ksw2 are adsorption and
⎛ ⎛ ⎞ ⎞
⎜ ∇p ⎟ 1 desorption rate parameters and Qswmax is the maximum
λ 4 ,k = λ40,k ⎜1 + ⎜ ⎟
⎟⎟ −6 ,

⎜ ⎝ λ41,k ⎟ α
10 + C se42,k ( 42 ) swelling state.
⎝ ⎠ ⎠ The effect of the swelling is that the pore volume
k=s,p. available for flow will be reduced. We model this by
reducing the effective water phase saturation with an amount
The last salinity term is optional, and requires the equal to the trapped water. Consequently, the relative
presence of at least one cation in the simulations. Cse is permeability to water will decrease during swelling. The oil
defined as the equivalent concentration of a reference cation, relative permeability will be affected indirectly through a
normally sodium. ∇p in the above expression is corrected for decrease in maximum oil saturation equal to trapped water.
the effect of pore throat plugging, i.e., this is subtracted from The amount of trapped water is corrected for the initial
the actual pressure gradient. content of water in the clay.

Effective salinity Results


Both the state of swelling clays and the attachment of This section presents simulations of various static and
fines to the pore surface are strongly dependent on the type dynamic mud filtration experiments and a dilute particle
of exchangeable cations such as K+, Na+, Mg2+, Ca2+, and on transport experiment, and demonstrates simulations of fines
the total salinity. The equilibrium between cations in solution migration and clay swelling. Finally, laboratory results are
and those attached to the clay surface is approximated by the used in simulation of a North Sea well. Mud compositions
mass action equations (Hirasaki 1982; Hirasaki and Lawson are given in Table 1.
1982): Input parameters used in the simulations (if not
z z mentioned in the text) are given in Table 2 (polymers), Table
f Nai X Nai 3 (solids), Table 4 (various model parameters), Table 5 (pore
= K XCi , ( 43 ) throat trapping), Table 6 (pore lining adsorption), Table 7
f Xi Xi (relative permeability) and Table 8 (dynamic filtration).
where Xi is the bulk phase concentration of cation i in N
(normality or meq/ml), zi is the valence and fXi is the fraction Static filtration
of cation i at the clay surface. KXci is an ion exchange The filter cake model is demonstrated by computing static
constant for each cation. XNa and fNa represent the bulk phase HPHT filtration experiments with two muds performed at
concentration and the surface fraction, respectively, of the 85°C and 850 psi on 20 µm and 60 µm ceramic discs. The
reference cation sodium. solids size distributions for the two mud systems are given in
To evaluate the salinity effect in mixed brines on e.g., Figure 2. The manufacturer permeability of the two discs is 5
clay stability, we adapt the effective salinity model from and 20 Darcy, respectively. These permeabilities and a
Utchem Pope and Nelson 1978; Hirasaki 1982 and extend it porosity of 0.37 were used in the calculations. Filtration
to any number of cations. Sodium (Na+) is used as reference volumes are computed for a 45 cm² filtration area (API).
cation, and the effective strength of other cations is The 0.0722 volumetric concentration of solids in mud 1
transformed to an equivalent amount of Na+. The effective (S1 in Table 3) includes a mixture of sized CaCO3 and barite.
salinity, Cse, is calculated from the fractional concentrations Average polymer properties are arbitrary used to match the
at the clay surface (for n cations): filtration volume, see P1 in Table 2. The polymer
concentration was set to 0.025.
X Cl Mud 2 contained 0.0377 vol. frac. sized CaCO3 (S2 in
C se = n −1
. Table 3). The only other parameter changed in the mud 2
1 − ∑ β ci f Xi
( 44 ) calculations was the intrinsic polymer viscosity [η], which
was increased from 1500 to 2500 ml/g (P2 in Table 2). All
i =1
other parameters remained unchanged.
The total salinity is here represented by XCl (aqueous Details from computation of the spurt loss and filtration
chlorine concentration), which is equal to the sum of all volumes for mud 1 on the 5 Darcy filter are given in Table 9.
cations in solution. βci is the effective salinity parameter for The main steps include:
cation i. 1. Determine critical size of solids that can enter the
formation (αsDφ=0.4·35.8=14.3 µm), Eqs. 6 and 22.
Clay swelling 2. Use size distribution to find concentration and specific
The net transport of water into or out from a swelling clay surface area of particles trapped at the disc surface.
is driven by osmotic forces, a gradient in the effective Compute average deposition thickness needed to form
salinity. The swelling is modelled as a kinetic water the initial cake, hsp = nsp6/S0sp = 21.1 µm, Eq. 23.
adsorption model with Cse the driving force in both terms: 3. Compute spurt loss volume, Vsp = 1.25 ml, Eq. 24.
SPE 122241 9

4. Determine average filter cake kca=9.3·10-6 mD and filtrate volume is only through the change in the brine
<φc>=0.416 from average cake properties <φo>=0.515, viscosity, from 0.926 in the API test (25°C) down to 0.318 cp
<nc>=0.0244 bar-1 and Δp=58.6 bar, Eqs. 11, 12 and 26. in the core tests (90°C). The effect of composition is more
5. Finally, compute the static filtration parameter complex. In the core experiments we observe a reduction in
Cst=0.0309 ml/cm²s0.5 and fluid loss Vf(30 min) = 8.86 Vf when OCMA-clay is added (FD1 versus FD2). In the
ml from Eq. 4. calculations, the filter cake in FD2 will be thicker, but with a
Experimental and simulated filtrate volumes are plotted higher kca because of the reduced relative polymer content.
versus square root of time in Figure 3 (mud 1) and Figure 4 However, adding OCMA decreases the average solids
(mud 2). We observe that increasing the permeability (pore particle size which may potentially trap more polymer in the
size) of the discs results in a higher Vsp. This is because a cake. In the simulations, approximately 9% of the xanthan
smaller fraction of the solids will be held back at the disc and 30% of the starch is passing through the filter cake in
surface, and in addition the average size of retained particles FD1. In FD2, these numbers are reduced to zero for xanthan
will be larger, so that a larger volume is required to form the and 1% starch. The increased entrapment of polymer in the
initial cake. Vsp is more than doubled for mud 2 compared to filter cake will reduce kca and this is the reason, at least in the
mud 1. This is mainly because of the lower solid simulations, for the observed decrease in Vf between FD1 (12
concentration. All these differences are well reproduced in ml) and FD2 (10 ml).
the calculated curves as clearly illustrated in Figure 3 and 4. The relative viscosity of the filtrate produced from mud
The second observation is that the filtration curves are 3a was measured to be approximately 6, and showed shear
parallel to each other, i.e., the filtration parameter Cst does thinning properties. This verifies that a significant amount of
not depend on the properties of the formation, but only on the polymer was passing through the filter cake in FD1. The
mud properties and the differential pressure over the cake. lower level of simulated polymer migration through the cake
Figure 5 shows computed filtrate volumes with mud 1 for in FD2 agrees well with measurements done by Sánchez et
a wide range of permeabilities. We show both the al. (2003), where 1-2% of the polymer was measured (by
analytically computed Vf and Vsp ignoring the pressure drop TOC) to pass through the cake for a mud system similar to
over the disc, and simulated Vf (30 min) using a 10 cm core. mud 3b. However, the actual values of the model parameters
We observe that Vsp starts to increase above 10 mD, while the used to simulate this phenomenon are highly uncertain. In the
net filtration volume after the spurt loss (Vf-Vsp) is constant in simulation we identify the degree of polymer entrapment in
the analytical calculations. The simulated Vf deviates from the cake as a potential important variable for filtration rates,
the computed value at low permeability (0.01 mD) when the however, the properties used in these calculations are poorly
resistance over the core becomes significant compared to the known. The polymer molecular weight Mw is well known as a
filter cake resistance. Also at very high permeability (100 physical property, but its actual value and its size distribution
Darcy) the simulated Vf is lower than the computed one, are highly uncertain. The computation of molecular hydraulic
because of internal blockage of pores that reduces the size from [η] and the trapping parameter αp are also
effective pore size of the formation. uncertain. Finally, the OCMA properties used are also at best
approximate.
Blaxter sandstone Return permeability. The oil return permeabilities in
The results from three static filtration experiments using Table 10 are obtained after back-flooding the core with oil at
Blaxter sandstone are given in Table 10. The core dimensions increasing rates, 1, 2.5, 5 and 8 ml/min, 30 min at each rate.
were 8 cm length and 2'' diameter with an effective filtration Simulated and experimental return permeabilities are in the
area of 16.8 cm2. The permeability was approximately 150 same range, 80-90% of initial oil permeability koi. The
mD and the porosity 20%. The mud system used in core simulated differential pressure over the core and the
experiment FD1 contained 50 g/l sized CaCO3 with size development of the return permeability RKO during the back
distribution given in Figure 2 (Mud 3a), 4 g/l xanthan and 20 flood of FD1 are shown in Figure 7. We see that RKO is
g/l low reticulated starch. 40 g/l drilling solids (OCMA clay) highly dependent on the length of the back flooding period.
was added in experiment FD2 and FD3 (Mud 3b). All the The profile is governed by relative permeabilities, which are
core floods were conducted at 90°C. An overbalance pressure unknown, and viscosity ratio between oil and water, which
of 15 bar was used in FD1 and FD2, and 35 bar in FD3. API may be reduced because of polymer invasion. The relative
filtration (100 psi and 20°C) and HTHP (500 psi and 90°C) permeabilities used (see Table 7) are typical for water-wet
are also reported in Table 10. medium, with a final tuning to the results in FD1, except for
Properties of the polymers and solids are given in Table 2 the oil end-point permeability.
and Table 3. Figure 6 shows a good agreement between The average Sw shown in Figure 7 is still well beyond its
measured and computed shear thinning mud viscosity. Other initial value 0.2. Also shown is RKOE, an estimated return
model parameters used in the simulations are given in Table permeability at maximum oil saturation. If the flood was
4, Table 5 and Table 6. continued, RKO (0.824) should finally reach RKOE (0.916).
Filtration. All the experimental fluid loss volumes in Note that capillary pressure is ignored (zero) in these
Table 10 are well reproduced in the simulations. This simulations and that additional uncertainty in return
indicates that the effects of pressure, temperature and permeabilities may be caused by capillary end effects.
composition are all well taken care of in the simulations. The Damage profile. Another issue much discussed is how
pressure sensitivity (7-35 bar) is handled by the compressible the damage profile looks like in the invaded zones. The
filter cake model. The simulated effect of temperature on simulated damage profile (RKO) for FD1 is shown in Figure
10 SPE 122241

8 together with profiles of trapped CaCO3 and polymers. upscaling a short dynamic experiment to the field situation,
CaCO3 particles are trapped in the first mms of the core, which will be shown in a later section.
while polymer shows a deep penetration. It seems that The second dynamic case in Figure 14 is a filtration
trapped polymer has reached a plateau extending about half experiment taken from Dewan and Chenevert (2001). A
the filtrate invasion volume. The damage profile (RKO) bentonite mud with solid content csm= 0.18 is used.
follows the trapped polymer profile. The amount of external Appropriate rheologic mud properties were simulated using a
CaCO3 particles introduced during the spurt loss period is polymer component. The solid size distribution was taken
obviously very small, and there is no realistic way to from one of the other muds (Duratherm) used in their paper.
simulate any serious damage from trapped CaCO3 particles The simulated slowness curve in Figure 14 matches the
alone. The observed damage is therefore most likely caused experimental curve very well. The experiment starts with
by the polymer. dynamic filtration (100 s-1) at 500 psi, switches to static
Another experiment FD8 shows very serious plugging conditions and back to dynamic conditions again, and ends
when only the polymer mixture is injected (4 g/l xanthan+20 with increasing the pressure to 1500 psi. Each period was 30
g/l starch) shown in Figure 9. When simulating FD8 with min, which was insufficient to reach steady-state conditions.
polymer trapping parameters given in Table 5, only the first After the pressure increase, we first see a short transient
part of the pressure drop curve was matched. The simulated period where the compressible cake adapts to the new
pressure shown in Figure 9 was obtained using a polymer pressure. Then the curve will slowly progress towards the
modification factor fpm=0.15. We are primary interested in same steady-state limit as before the pressure increase,
the effect of much smaller polymer amounts and find that although at a slower rate because more mass must
using fpm=1 (as in Table 5) is fine. accumulate in the cake at the higher pressure.
In Figure 10, the back flooding of core B14 is extended The uncertainties in the particle size distribution and the
with an extra step where the "final" oil in place is displaced rheologic properties make the value of the model parameters
with a high viscous oil (M82) at reduced rate. Using the in Table 8 less general. The main purpose was, however, to
pressure derivative with respect to the front position of M82, demonstrate the capability of the dynamic model to
the permeability profile along the core can be estimated. We reproduce typical experimental behaviour.
approximate the front position by assuming constant So
throughout the core. In Figure 11 we show the estimated ko Particle retention
profile (Exp. VF) together with initial koi. We also show An experiment (Exp. 1) taken from Al-Abduwani et al.
profiles from simulation of the experiment and an estimated (2003) is used to demonstrate the pore throat trapping model.
ko-profile from the simulated pressure history (Sim VF). Note In this experiment, a dilute 20 ppm hematite (Fe2O3) particle
that the estimated increase in ko towards the outlet end in the suspension was injected into a Bentheimer sandstone core.
oil flood, which is to the left in Figure 11 and the mud side, is The particle size was stated to be in the range 0.1-5 µm with
a result of dispersion of the viscous oil front. The same 65% less than 1 µm. The size distribution used in our
behavior was seen when applying the method on simulated simulations is given in Figure 15 and additional information
data (Sim VF). in Table 3. Core data used in the simulations are L=12.2 cm,
B14 has a length of 4.7 cm, and was subjected to static k=1300 mD, φ=0.22 and Sw=1. The trapping parameters for
mud exposure for 50 hours (mud 3b), during which close to 2 solids tuned to this experiment (Table 5) are also used in the
pv filtrate had entered the core. The results in Figure 9 other simulations presented in this paper.
strongly indicate that ko is reduced throughout the whole By using dx/dt=uw/(φSw) in Eq. 36, setting λ2 to zero and
core, which supports the previous results (Figure 8) integrating, we find the effluent concentration for a given
indicating that polymers may cause deep, and rather evenly particle size to be:
distributed damage.
c = c 0 exp(− λ1 L ) , ( 46 )
Dynamic filtration where λ1(Dp) is given by Eq. 37. The computed size
Two examples of simulating dynamic experiments are dependent relative effluent concentration for the present core
given in Figure 12 to Figure 14. The first example is a core is shown in the size distribution plot (Figure 15). The average
test FD15 (Blaxter sandstone) with mud 3b (described effluent concentration is estimated from the size distribution
earlier). The experiment is simulated with three different sets to be 13.1 ppm when 20.7 ppm is injected.
of parameters (fn, nF, β, τy) given in Table 8. The plot of Vf Simulated and experimental results are compared in
versus time in Figure 12 has a non-uniform development of Figure 16 to Figure 19. If pore throat trapping is considered
its slope which may suggest some experimental problems. as the only retention mechanism, the simulated effluent
This is more clearly seen when plotting the slowness (inverse concentration seen in Figure 16 starts at the 13 ppm
velocity) versus Vf as done in Figure 13. Simulation DS3 puts analytical value and then declines slowly as an effect of
more weight on the last part and assumes a steady-state limit, previously trapped material. The simulation was extended
of approximately 6000 sec/cm, is reached. DS4 puts more beyond the experimental time period, and at some point,
weight on the first part, while DS2 is in between. DS2 and when an external filter cake has been formed, the effluent
DS4 have not reached steady-state within the 4 hours concentration dropped rapidly to zero. The experimental
timeframe. In DS2 no erosion at all is assumed, only a curve is different, it breaks through at a low concentration
reduced attachment rate. This illustrates the problem of and increases rather slowly towards a plateau after
SPE 122241 11

approximately 300 pv. This curve shape may be in part passing caused by the heterogeneous permeability
explained by a gradual change in the trapping pattern due to a distribution.
distribution in pore throat size, but it is also influenced by Figure 23 and Figure 24 show the computed permeability
other retention mechanisms. A closer match to the impairment along the well. The normalized oil permeability
experimental profile is obtained by including pore-lining for each zone is computed by two methods in Figure 23, (a)
adsorption in the experiment. as a productivity index computed from the rate in each zone
A total of 662 pv was injected in Exp. 1. The final and (b) analytically ignoring cross-flow between zones
distribution of hematite and permeability along the core is (RKO). The error introduced by ignoring cross-flow is very
shown in Figure 17 and Figure 18, respectively. In both small in this case, and almost zero for the total well with
cases, the simulated and experimental profiles match. The relative PIoil and RKO equal to 0.794 and 0.792, respectively,
experimental data represents averages over core segments. A in FS4. The same results are plotted in terms of skin
corresponding segmental permeability distribution computed distribution in Figure 24 where S (current) corresponds to
from the simulation results is almost overlapping the RKO. S (kabs) is computed from the reduction in absolute
experimental curve shown in Figure 18. permeability, corresponding approximately to RKOE. The
Figure 19 shows the specific surface area of trapped radial distribution of RKO and RKOE along the well is given
hematite particles throughout the core from the simulation. A in Figure 25.
corresponding effective "average" particle diameter We conclude from the upscaling of laboratory results
computed as Dp=6φsh/S0 is also shown. Obviously, average with mud 3b, that no significant real damage to the formation
size of trapped particles is larger at the inlet end, while should be expected from externally introduced particles in
smaller particles will penetrate deeper. the present case. Other internal damage processes are not
evaluated in this example. The "real" damage as indicated by
Fines migration and clay swelling RKOE in Figure 25 is in line with the observed polymer
Simulations of fines migration and clay swelling upon damage observed in the laboratory experiments. One should,
injection of low salinity brine are demonstrated in Figure 20. however, expect a temporary reduced productivity because of
In the fines simulation, 1% of the rock matrix is defined as incomplete displacement of mud filtrate. The size and
fines, held in place using the pore lining adsorption model. In duration of this decline depend on filtrate penetration depth,
the clay swelling simulations, 2% of the rock matrix is relative permeability and the variance of the permeability
defined as swelling clay with a maximum swelling capacity distribution.
of 10 (Qswmax in Eq. 45). Other parameters in Eq. 45 are The fluid loss volumes in Table 11 are very high, in fact
ksw1=0.0002 meq/(ml·min) and ksw2=0.001 meq/(ml·min). The an order of magnitude larger than expected. The worst case is
effective salinity, Cse, is essential equal to the NaCl FS3, in which a high erosion rate is assumed. Using a steady-
concentration indicated in Figure 20 and 0.7 meq/ml with the state slowness of 5628 sec/cm corresponding to that obtained
initial brine. for DS3 in Figure 13, but corrected for the different shear
rate in FS3, and the filtration area of the well (πDwLw = 1094
Field scale simulations m²), we compute a filtrate volume of 1750 m³ compared to
To test the upscaling to field conditions, we use mud 3b 1717 m³ given in Table 11. Similarly, the static fluid loss for
on a horizontal well from the North Sea. Permeability and the well is estimated to 55.6 m³ using the filtration parameter
porosity distribution along the wellbore is given in Figure 21. Cst from FD3 (0.0032 m³/m²hr0.5), which compares well to
We set the wellbore diameter to 0.24 m, and use a constant the simulated 55 m³ in Table 11. The deviations are because
pressure boundary at 100 m in our 2D radial model. The of differences in overbalance pressure and low permeable
temperature is 90°C. Filtrate invasion is simulated for a zones offering flow resistance comparable to the mudcake
period of 250 hours with a overbalance pressure of 86 bar. resistance. These computations indicate that the field scale
Then we simulate cleanup by back producing the well using a simulations are correct in terms of upscaling the laboratory
30 bar drawdown for 24 hours. Four cases are run, FS1 with results, but our description of the field conditions may be
static filtration and FS2-FS4 with dynamic conditions at insufficient. Of course, our interpretation of the laboratory
shear rate of 170 s-1. Cases FS2-FS4 correspond to various results is very important, and in the case of FD15, the
matches to the dynamic core experiment FD15, and dynamic estimated dynamic parameters are very uncertain. But all the
filtration simulations DS2-DS4 shown in Table 8. interpretations used predicted large fluid loss volumes in the
The simulations are summarized by main numbers in field case. Other explanations should be sought.
Table 11. Typical results from the simulations are One obvious thing is that the formation will be gradually
demonstrated by simulation FS4 in Figure 22 to Figure 25. exposed to the mud as the well is drilled. In the simulations,
Figure 22 shows the saturation distribution around the well the total well was exposed from time zero. Exposing the
after mud exposure for 200 hours and after the 24 hours back formation according to a well penetration rate of 10 m/hr is
production period. The variation in penetration depth is shown in Figure 26 (FS4b). This will, however, only result in
mainly a result of the porosity distribution along the wellbore a moderate reduction of the filtrate volume from 702 to 535
because filtration rate through the filter cake is independent m³. Simulating with periods of 50 hrs dynamic + 25 hrs
on the formation properties. After the back production, we static, and last 50 hrs static, reduces the loss to 251 m³
see that there is still significant amount of water around the (FS4c). In simulation FS2, only reduced mudcake growth and
well, indicating that substantial production is needed to re- no erosion are assumed. Applying the same periodic
establish original fluids saturation. This is a result of by- dynamic/static history on FS2 reduces the fluid loss from 504
12 SPE 122241

to 225 m³ (FS2c). Variations in shear rate are obviously Nomenclature


important, and knowledge about how the mudcake will a Mw exponent for polymer viscosity
respond to increased shear stress after a static period is vital A area, [L2]
for good predictions at the field scale. Other possibilities that A adsorbed material [pv], solids or polymer
may contribute to higher than expected computed fluid loss Bp viscosity parameter for pressure, [Lt2/m]
at the field scale are that the effective shear rate is lower than BT viscosity parameter for temperature, [T]
projected because of off-centered drill string and the presence BvT polymer viscosity parameter, [T-1]
of significant surface roughness in the well compared to the c concentration (volume fraction)
flat core ends used in laboratory experiments. C1 filtration coefficient, [L³/L³]
Cse effective salinity, [N]
Summary and conclusions Cst static filtration parameter, [L3t-0.5]
Maximize is a program for computation of formation damage D diameter, [L]
imposed on the formation during drilling. Among the main Dch polymer chain diameter, [L]
parts modelled in the program are filter cake build-up under Dhp hydraulic polymer diameter (polymer size in
static and dynamic conditions, fluid loss to the formation, solution), [L]
transport of solids and polymers in the formation including fn friction factor
effects of pore lining retention and pore throat plugging, and fX fractional concentration of cation at the rock
salinity effects on fines stability and clay swelling. The surface, [N/N]
models handle multi-component water-based mud systems at h thickness, [L]
both core and field scales. k permeability, [L²]
Our filter cake model differs from previous models in that k', k'' polymer viscosity parameters, Eq. 15
we compute the filter cake's porosity and permeability from ksw1 clay swelling parameters, [Nt-1]
properties ascribed to the individual solid and polymer ksw2 clay de-swelling parameters, [N-1t-1]
components in the mud. The necessary data for each KXC cation exchange parameter, [Nz-1] where z is the
component are porosity and compressibility, if packed as a valence
"sand pack", and specific surface area. The specific surface Mw molecular weight, [g/mol]
area is obtained from the size distribution for solid particles N normality, [meq/ml]
and from the average chain diameter of polymers. The n number, multiplayer
model's capability to reproduce various fluid loss nc filter cake compressibility, [Lt2/m]
experiments is demonstrated. For a CaCO3 mud, the effects nF dynamic filtration exponent
of varying temperature, differential pressure and adding p pressure, [m/Lt2]
drilling solids are all simulated with a single set of input Pα polymer shear thinning exponent, Eq. ( 17 )
parameters. pv pore volume, used as normalized volume unit
The particle retention is split into two main mechanisms, q flow rate, [L3/t]
pore throat trapping and pore lining retention. We focus RKA absolute permeability normalized on initial k
mainly on the pore throat trapping which is the process with RKO oil permeability normalized on initial koi
the main damage potential. A dilute particle injection RKOE ko at current maximum So normalized on initial
experiment was selected for testing the model. The ko(maximum So)
simulation provided excellent matching of final experimental S skin factor (dimensionless), saturation
profiles of retarded particles and permeability along the core. S0 specific surface area (particle surface/particle
Upscaling of laboratory result to the field scale is volume), [L-1]
demonstrated for a North Sea well. Only damage from t time [t]
externally introduced mud particles and polymers were T temperature, [T]
considered. The simulations indicated essential no permanent u Darcy velocity, u=q/A [L/t]
damage, in line with the moderate polymer damage observed V volume, [L3]
in the laboratory experiments. However, a temporary decline X ion concentration, [N]
in productivity due to slow displacement of invaded fluids Y erosion rate, [L/t]
was observed. The slow restoration of initial saturation z 1-φ, ion valence
around the well is attributed to the heterogeneous
permeability distribution and to large fluid loss volumes in α effective mud fraction in dynamic filtration
the simulations. The simulated fluid loss volume is an order α critical size parameter for particle flow in porous
of magnitude larger than normally observed in the studied medium
field cases. The dynamic fluid loss had been tuned to a αnk n=11, 22: trapping parameters for solids or
dynamic laboratory experiment, and analytical calculations polymer (k=s,p).
verify that the simulated results are correct in terms of n=41, 42: adsorption parameters for solids or
upscaling the fluid loss rate at the laboratory scale. polymers (k=s,p).
βc effective salinity parameter
βe erosion rate parameter, [t-1]
γ shear rate, [t-1]
SPE 122241 13

γhf polymer viscosity parameter, [t-1] SPE 17617 presented at the SPE International Meeting on
Petroleum Engineering, Tianjin, China.
[η] 3
intrinsic polymer viscosity, [L /m] Aziz, K. and Settari, A. (1979). Petroleum Reservoir Simulation.
λnk n=1, 2, 10, 20, 21, 22: trapping parameters for London and New York, Elsevier Applied Science Publishers.
solids or polymer (k=s, p), [L-1] Bezemer, C. and Havenaar, I. (1966). Filtration Behaviour of
n=3, 4, 40, 41: adsorption parameters for solids or Circulating Drilling Fluids. SPEJ 6 (4): 292-298.
polymers (k=s,p), [L-1] Bourgoyne, A. T., Chenevert, M. E., Millhelm, K. K. and Young, F.
S. (1986). Applied Drilling Engineering, SPE.
μ viscosity [m/Lt] Cerasi, P., Ladva, H. K., Bradbury, A. J. and Soga, K. (2001).
ν dimensionless intrinsic viscosity Measurements of the Mechanical Properties of Filtercakes.
σ trapped material, [pv] SPE 68948 presented at the SPE European Formation Damage
τ tortuosity parameter, (path length/length)2 Conference, The Hauge, The Netherlands.
τy yield strength (filter cake), [m/Lt2] Civan, F. (2000). Reservoir Formation Damage : Fundamentals,
Modelling, Assessment, and Mitigation. Houston, Texas, Gulf
φ porosity
Publishing Company.
Dewan, J. T. and Chenevert, M. E. (2001). A Model for Filtration of
Subscripts Water-base Mud During Drilling: Determination of Mudcake
a average Parameters. Petrophysics 42 (3): 237-250.
c cake (filter cake) Ding, Y., Longeron, D., Renard, G. and Audibert, A. (2002).
cr critical Modelling of Both Near-Wellbore Damage and Natural
d dynamic Cleanup of Horizontal Wells Drilled With a Water-Based Mud.
e end point SPE 68951 presented at the SPE International Symposium and
m mud Exhibition on Formation Damage Control, Lafayette, Lousiana.
Ferguson, C. K. and Klotz, J. A. (1954). Filtration from Mud During
o oil
Drilling. JPT (February): 30-42.
p particle, polymer Fordham, E. J., Allen, D. F. and Ladva, H. K. J. (1991). The
r relative Principle of a Critical Invasion Rate and Its Implications for
s solids Log Interpretation. SPE 22539 presented at the SPE Annual
sp spurt loss Technical Conference and Exhibition, Dallas, Texas, U.S.A.
sw water in swelling clay Gruesbeck, C. and Collins, R. E. (1982). Entrainment and
t trapping Deposition of Fine Particles in Porous Media. SPEJ 22 (6):
w water 847-855.
φ pore property Hirasaki, G. J. (1982). Ion Exchange with Clays in the Presence of
Surfactant. SPEJ 22 (2).
Hirasaki, G. J. and Lawson, J. B. (1982). An Electrostatic Approach
Acknowledgements to the Association of Sodium and Calcium with Surfactant
The major part of this work was done within the Well Micelles. SPE 10921.
Productivity 2002 project supported by the European Union Lake, L. W. (1989). Enhanced Oil Recovery, Prentice-Hall, Inc.
(EU) 5th Frame programme. The authors would like to thank Meter, D. M. and Bird, R. B. (1964). Tube Flow of Non-Newtonian
EU and the industrial sponsors of the project, Agip BP Polymer Solutions: Part I. Laminar Flow and Rheological
Exploration, ChevronTexaco, Norsk Hydro and Shell Models. A.I.Ch.E 10 (6): 878-881.
Norway. We would also like to thank the other partners in the Outmans, H. D. (1963). Mechanics of Static and Dynamic Filtration
project: ResLab, Eni Tecnologi, Rhodia and M-I Drilling In the Borehole. SPEJ (September): 236-244.
Pope, G. A. and Nelson, R. C. (1978). A Chemical Flooding
Fluids. Experimental data used in this paper is mainly Compositional Simulator. SPEJ (Oct.): 339-354.
supplied by ResLab and Eni Tecnologi. Sánchez, E., Audibert-Hayet, A. and Rousseau, L. (2003). Influence
The second phase of this work was supported by Shell, of drill-in fluids composition on formation damage. SPE 82274
BP and StatoilHydro. presented at the SPE European Formation Damage Conference,
Supply of lab test data for model calibrations by The Hague, The Netherlands.
Halliburton is much appreciated. Semmelbeck, M. E., Dewan, J. T. and Holditch, S. A. (1995).
Invasion-Based Method For Estimating Permeability From
Logs. SPE 30581 presented at the SPE Annual Technical
References Conference and Exhibition, Dallas, U.S.A.
Al-Abduwani, F. A. H., Hime, G. and Alvarez, A. (2005). New Sharma, M. M. and Yortsos, Y. C. (1986). Permeability Impairment
Experimental and Modelling Approach for the Quantification Due to Fines Migration in Sandstones. SPE 14819 presented at
of Internal Filtration. SPE 94634 presented at the SPE the SPE Symposium on Formation Damage Control, Lafayette,
European Formation Damage Conference, Scheveningen, The LA.
Netherlands. Sorbie, K. S. (1991). Polymer-Improved Oil Recovery, Blackie and
Al-Abduwani, F. A. H., Shirzadi, A., van der Broek, W. M. G. T. Son Ltd.
and Currie, P. K. (2003). Formation Damage vs. Solid Particles Suryanarayana, P. V., Wu, Z., Ramalho, J. and Himes, R. (2005).
Deposition Profile during Laboratory Simulated PWRI. SPE Dynamic Modelling of Invasion Damage and Impact on
82235 presented at the SPE European Formation Damage Production in Horizontal Wells. SPE 95861 presented at the
Conference, The Hague, The Netherlands. SPE Annual Technical Conference and Exhibition, Dallas,
Arthur, K. G. and Peden, J. M. (1988). The Evaluation of Drilling Texas, U.S.A.
Fluid Filter Cake Properties and Their Influence on Fluid Loss. Wennberg, K. E. and Sharma, M. M. (1997). Determination of the
Filtration Coefficient and the Transition Time for Water
14 SPE 122241

Injection. SPE 38181 presented at the SPE European The denominator is given in A.3 and the numerator integral
Formation Damage Conference, The Hague, The Netherlands. is after using the previous substitution, z=1-φ:
Wu, J., Torres-Verdin, C., Sepehrnoori, K. and Delshad, M. (2004).
Numerical Simulation of Mud-Filtrate Invasion in Deviated Δp z ( Δp )
−1 ⎡1 2⎤
Wells. SPE Reservoir Evaluation & Engineering 7 (2): 143- ∫ φc kdp n = 2n τS 2 ⎢⎣ z + 3 ln z − 3z + 12 z ⎥
⎦ z (0)
. (A. 8)
154. 0 c 0

By using
1
Appendix Fφ ( pn ) = + 3 ln z − 3z + 12 z 2 , (A. 9)
Computation of average filter cake properties z
The compressible cake permeability, kc(pn) can be we get the average porosity
Fφ (Δp ) − Fφ (0)
expressed using Eqs. 5 and 10:
φ = .
φo3e −3nc p n
(A. 10)
Fk (Δp ) − Fk (0)
k c ( pn ) =
( )S
2
. (A.1)
2τ 1 − φo e − nc p n 2
0

Substituting Eq. A.1 into Eq. 8 yields: Table 1: Mud compositions (g/liter).
Δp Δp Component Mud 1 Mud 2 Mud 3a Mud 3b
1 φo3e −3nc p n
∫ kc dpn = 2τS 2 ∫
Polymer (P1,P2) 25 25
dp n .
(1 − φ e )
(A.2)
− nc p n 2 Xanthan 4 4
0 0 0 o Starch 20 20 20
CaCO3 100 100 50 50
Defining z = 1 − φ = 1 − e − nc pn in Eq. A.2 yields: Drilling solids (OCMA) 40
Δp Barite 155
1
z2
(1 − z )2 dz,
∫ k c dp n = 2τS 2 n ∫ z2 Table 2: Polymer properties; numbers in italic do not influence
0 0 c z1
(A.3) static fluid loss computation.
Δp z ( p = Δp )
−1 ⎡1 ⎤ 2
∫ k c dp n = 2τS 2 n
Name P1 P2 Xanthan Starch
⎢ z + 2 ln z − z ⎥ .
0 0 c ⎣ ⎦ z1 ( p =0) Viscosity, low shear
[η] (ml/g) 1500 3200 5500 600
Define:
k' 0.4 0.4 0.4 0.4
1
Fk ( p n ) = + 2 ln z − z .
k'' 0.01 0.01 0.07 0.07
(A. 4)
z Mw exponent, a 0.8 0.8 0.8 0.5
BvT [1/°C] 0 0 0.006 0.001
The average filter cake permeability is then:
T0 (°C) 20 20 20 20
1 Fk (0) − Fk (Δp ) Shear thinning
k ca = ⋅ . (A. 5)
2τS 02 nc Δp Pα 1.7 1.7 1.8 1.5
γhf (s-1) 2 2 1.5 5
Let kca(x)=kca(pn(x)) be the average permeability from the Size parameters:
cake surface to depth x within the cake. Because the rate and 7 7 6 6
Mw (g/mol) 10 10 3·10 5·10
viscosity of the permeate flow are constant, the product
σ ln(Mw) 1 1 2 2
kca(pn(x))·pn(x)/x must be constant. If we define x as the
relative position within the cake as a function of pn then the Dch (µm) 0.0008 0.0008 0.001 0.0012
distance from the cake surface to pressure pn(x) is: ρ (g/cm³) 1 1 1 1

Δhc Fk ( p n ) − Fk (0)
Computed properties:
x( p n ) = = . (A. 6) S0 (m²/cm³) 5000 5000 4000 3333
hc Fk (Δp ) − Fk (0) Mean Dhp (µm) 0.414 0.498 0.428 0.242
In order to calculate the flow rate, we also need to know the Filtercake properties:
thickness (dc) of the cake. This is readily obtained from mass φo 0.85 0.85 0.85 0.85
balance once the average porosity of the cake is found. Using nc (bar )
-1
0.095 0.095 0.08 0.08
Eqs. 7 and 8 we can write the average porosity as:
Δp

1 ∫ φk c dp n
φ = ∫ φ c dx = 0
. (A. 7)
Δp
∫ k c dp n
0

0
SPE 122241 15

Table 3: Solids properties. S0 is computed. Table 8: Dynamic filtration parameters.


CaCO3
*
OCMA Sim # fn nF βe (s-1) τy (Pa)
* * *
Name S1 S2 mud 3 mud 3b FeO3 DS2 0.1 1.25 0 5000
**
Dp (µm) 18.44 29.4 7.05 1.5 0.856 DS3 0.1 0.5 67 5000
***
St Dev 1.4 1.34 0.8 1 0.95 DS4 0.05 0.8 1 1000
φsh 1.8 1.35 1.25 1.5 1.8 Mud98067 300 0.5 1 10
S0 (m²/ml) 1.5 0.67 1.5 9.91 19.6
Table 9: Computation of static filtration, mud 1, Ac=45 cm².
ρ (g/cm³) 2.6 2.6 2.7 2.7 5.4
Model αs αp nsp [η] μs
Filtercake:
parameters 0.4 0.1 1 1500 0.339
φo 0.4 0.4 0.4 0.4 0.4
-1
Disc K φ Ac Dφ αs·Dφ
nc (bar ) 0.0001 0.0001 0.0001 0.001 0.0005 properties (mD) (cm²) (µm) (µm)
*) Tabular distribution, mean Dp and St. Dev. are computed. 5000 0.37 45 35.8 14.3
**) Dp corresponding to mean ln(Dp).
***) Standard deviation in ln(Dp). Spurtloss Fraction > S0sp hsp C1sp Vsp
αs·Dφ (m²/ml) (µm) (ml)
Table 4: Various model parameters, values used in simulations. 0.589 0.284 21.1 0.0763 1.25
Variable Value Remarks Mud Conc, Eff. Conc S0 φo nc
-1
(vol. frac.) (vol. frac.) (m²/ml) (bar )
αs 0.4 size factor in spurt loss
Polymer 0.0250 0.0249 5000 0.850 0.0950
αp 0.1 size factor for polymers
Solids 0.0722 0.0722 1.5 0.400 0.0001
nsp 1 spurt loss factor
Total 0.0972 0.0971 1283 0.515 0.0244
τ 3 tortuosity factor
Filtration <φc > <kc> C1 Cst Vf 30 min
0.5
Table 5: Pore throat trapping parameters. Δp=58.6 bar mD ml/cm²s (ml)

Solids Polymers 0.416 9.27E-06 0.1993 0.0309 8.86

αt (fraction) 0.035 0.05


Table 10: Experimental and simulated fluid loss volumes with
λ10 (cm-1) 40 1.5 mud 3.
α11 2 2.5 FD1 FD2 FD3
λ20 (cm-1) 0.0001 0.0015 Δp (bar) 15 15 35
λ22 (bar/cm) 0.01 Drill solids (g/liter) 0 40 40
α22 0.5 API (100psi, 20°C) exp 6 5 -
fpm 1 Vf - 30min (ml) sim 5.9 4.7 -
fpt 0.05 *
HTHP (500psi, 90°C) exp 12 12 -
fpt2 0.1 Vf - 30min (ml) sim 12.4 10.6 -
**
Core (90°C) exp 12 10 11
Table 6: Pore lining adsorption parameters.
Vf - 240min (ml) sim 12.1 10.3 11.1
Solids Polymers Fe2OH3 Fines
Oil return exp 83 82 86
Qm [pv] 0.05 0.001 0.05 0.2
permeability % sim 82.9 86.8 85.4
lamda30 [1/min] 0.1 0.1 0.1 0.03 *) HPHT filtration volumes converted to API filtration area.
lamda31 0 0 0 0 **) Core filtration area is 16.8 cm2
lamda40 0.0001 0.0002 0.0001 0.0001
Table 11: Field case simulation result, after 250 hrs mud
lamda41 [bar/cm] 10 0.1 10 10 -1
exposure (86 bar, 170 s ) and after 24 hrs back production (30
alfa41 3 1 3 0.25 bar drawdown).
3
alfa42 1.5 1.5 1.5 1.5 Sim # Vf (m ) RKO S (current) S (abs k)
RPRmax 2 2 2 2 FS1 55 0.969 0.22 0.011
FS2 504 0.872 1.0 0.005
beta 200 1000 200 1000
FS2c 225 0.924 0.6 0.008
FS3 1717 0.732 2.3 0.003
Table 7: Relative permeability, Corey parameters.
FS4 702 0.842 1.2 0.004
Sim krwe Ew Srw kroe Eo Sor
FS4b 535 0.833 1.3 0.005
FD1-FD3 0.15 3 0.2 0.77 1.5 0.2 FS4c 251 0.895 0.8 0.008
FD15 0.15 3 0.2 0.65 1.5 0.2 FS4b: Included rate of penetration 10m/hr
FS2c, FS4c: Dynamic + static periods (50 hrs + 25 hrs), static the last 50 hrs
B14 0.1 4 0.185 0.75 3 0.185
Field 0.4 3 0.14 0.65 4 0.1
16 SPE 122241

0.7 16

0.65 1st curve 14


Firf=0.2
0.6 12
Firf=0.8
0.55
Porosity

10

Vf (ml)
0.5
8
0.45
6
0.4
4
0.35
n c =0.04
2 20 µm 3 60 µm 3
0.3
20 µm sim 60 µm sim
0 5 10 15 20 25
0
Cake overburden pressure (bar) 0 1 2 3 4 5 6

Figure 1: Compressible filter cake porosity, calculated with Sqrt (min)


φo=0.65, nc=0.04 bar-1 and irreversibility factor Firf=0.2 and 0.8. Figure 4: HTHP filtration at 85°C and 850 psi, experimental and
simulated for 20 and 60 µm filter discs, Mud 2. (Ac=45cm²).
1
Cumulative distribution, F (vol.fraction)

100
Vsp 1 Vf-Vsp (1)
0.8
Vf 1 Vf sim (1)
Vsp 2

0.6
Filtrate volume (ml)

10

0.4

0.2 Mud 1 1
Mud 2
Mud 3
0
0.1 1 10 100 1000

Particle diameter (µm) 0.1


0.001 0.1 10 1000 100000
Figure 2: Particle size distribution.
K (mD)
Figure 5: Effect of the porous medium on filtrate volumes.
12 Analytical spurt loss (Vsp) and filtrate volume (Vf at 30 min) and
simulated with a 10 cm core (Vf-sim). (Ac=45cm²).
10
10000
Mud 3a
8 Calc
Exp.
Vf (ml)

1000
Viscosity (mPa·s)

4 100

2
20 µm 1 60 µm 1 10
20 µm sim 60 µm sim
0
0 1 2 3 4 5 6
1
Sqrt (min) 1 10 100 1000 10000
Figure 3: HTHP filtration at 85°C and 850 psi, experimental and Shear rate (1/sec)
simulated for 20 and 60 µm filter discs, Mud 1. (Ac=45cm²).
Figure 6: Mud 3a, experimental (converted Fann viscosities) and
calculated, 25°C.
SPE 122241 17

0.75 1 0.6 0.6

0.5 0.5
0.6 0.8

Oil return permeability, RKO


8 ml/min
Pressure drop (bar)

Pressure drop (bar)


Ko/Koi and <Sw>
0.4 Sim. 0.4
0.45 0.6
Exp. CLS370
5 ml/min
0.3 Exp. M82 0.3
0.3 0.4 RKO
2.5 ml/min
0.2 0.2
0.15 1 ml/min 0.2
dp RKO
0.1 0.1
RKOE <Sw>
0 0
0 20 40 60 80 100 120 0 0
Time (min) FD1 0 5 10 15 20 25 30 35 40

Figure 7: Simulated back production, FD1. Differential pressure PVinj B14: back production
(dp), current (RKO) and maximum (RKOE) oil return permeability,
and average Sw. Figure 10: Back production with low viscosity oil (CLS370) and
high viscous oil (M82) in experiment B14.

160
1 0.001
140
Trapped CaCO3×1 and

0.8 0.0008 120


polymer×100 [pv]

RKOE 100
0.6
Ko
0.0006
Ko/Koi

RKO 80
CaCO3pt 60
0.4 0.0004 Sim. Sim VF
xanthan 40
Exp. VF Kio
0.2 starch 0.0002 20
Ko max
0
0 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Core position, x/L B14
x/L
Figure 11: Interpretation of oil permeability profile (Exp. VF) from
Figure 8: Simulated oil permeability reduction, current (RKO) and
viscous oil flood, B14.
estimated final (RKOE), and pore throat trapped solids (CaCO3)
and polymers.
50
45 exp.
35
40 DS2
Sim
Filtrate volum (ml)

DS3
30 Exp. FD8 35
DS4
30
Pressure drop (bar)

25
25
20 20
15
15
10
10 5
FD15
0
5 0 50 100 150 200 250

0 Time (min)
0 1 2 3 4 5
Figure 12: Dynamic filtration Blaxter sandstone, Mud 3b.
PV injected Simulation of experiment FD15.
Figure 9: XC+Starch injection, 150 mD Blaxter sandstone.
18 SPE 122241

0.008
10000

FeO3 concentration (pv)


0.007 Sim
9000 exp.
0.006 Exp.
8000 DS2
0.005
Slowness (sec/cm)

DS3
7000 0.004
DS4
6000 0.003

5000 0.002
0.001
4000
0
3000 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
2000
1000 Figure 17: Experimental (Al-Abduwani et al. 2003) and simulated
FD15 hematite distribution in core after 662 pv injection.
0
0 10 20 30 40 50

Vf (ml)
1

Normalised permeability
Figure 13: Slowness vs. Vf in dynamic filtration experiment FD15.
0.8

0.6
60000
Exp. Mud 98067 0.4
Sim Sim
50000
0.2 Sim. segment
Slowness (sec/cm)

Exp. segment
40000 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L
30000
Figure 18: Experimental (Al-Abduwani et al. 2003) and simulated
20000 permeability profile in core after 662 pv injection.

10000 12 3
500 psi 500 psi 500 psi 1500 psi

Trapped particle D (µm)


100 sec-1 0 sec-1 100 sec-1 100 sec-1 10 2.5
S0 trapped FeO3

0
8 2
0 30 60 90 120
(m²/cm³)

6 1.5
Time (min)
4 1
Figure 14: Experimental, replotted from Dewan and Chenevert S0
(2001), and simulated dynamic filtration experiment. 2 0.5
Dp (µm).
0 0

1 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


x/L
0.8 0.8 Figure 19: Profiles of specific surface area and corresponding
Relative effluent conc.
Cum. vol. fraction

F Hematite
Ceff
particle size of hematite in core after 662 pv injection (simulated).
0.6 0.6
Ceff (Dlim)
0.4 0.4
1000
0.2 0.2
Fines
0 0 Swelling clay
0.1 1 10 100
Particle dimeter (µm)
Pressure drop (bar)

Figure 15: Hematite particle distribution used in simulation.


Relative effluent concentration is for the Bentheimer core
10
described in the text, computed with (Dlim) and without a lower
size limit for trapping (αts).

16 1
Effluent hematite (ppm)

Sim
14
Exp.1 0.5 N 0.1 N 0.01 N 0.001 N
12 Exp.2 Brine Brine
NaCl NaCl NaCl NaCl
10 Exp.4
Sim A=0 0.1
8 0 5 10 15
6 PV injected
4
Figure 20: Simulation of fines mobilization and clay swelling.
2
0
0 200 400 600 800 1000 1200 1400 1600 1800
PV injected

Figure 16: Experimental and simulated effluent profile for


hematite. Pore lining adsorption is turned of in "Sim A=0".
Experimental data is replotted from Al-Abduwani et al. (2003).
SPE 122241 19

#FS4 Oil RKO


1
Permeability Porosity 0.9
0 0 0.8
0.7

PI (relative)
lg(mD) fraction
0.6
2.686 0.248 0.5
0.4
200 200 0.3
0.2
0.1
1.765 0.196 0
0 200 400 600 800 1000 1200 1400
400 400 Wellbore (m)
0.845 0.145
Figure 23: Sim. FS4. Relative productivity index for oil, from
simulated well rate and computed ignoring crossflow (RKO). After
Wellbore (m)

600 Wellbore (m) 600 24 hrs production at 30 bar drawdown.


-0.075 0.093

#FS4 S (current) S (Kabs)


100
800 800
-0.996 0.041
10

Skin
1000 1000 0.1

0.01

0.001
1200 1200
0 200 400 600 800 1000 1200 1400
Wellbore (m)

Figure 24: Sim. FS4. Skin distribution computed from current


1400 1400
productivity and from current absolute permeability. After 24 hrs
production at 30 bar drawdown.
0.0 0.5 1.0 1.5 2.0 0.0 1.0 2.0
Radial direction (m) Radial direction (m)
RKO, #FS4, 24hrs RKOE, #FS4, 24hrs
Figure 21: Field case, permeability (logarithmic) and porosity back production back production
distributions. 0 0
RKO RKOE
1.000 1.000
So, 24 hrs. Back
So, 200 hrs. #FS4. production. #FS4 200 200
0 0
So So 0.750 0.950
0.860 0.860
400 400
200 200
0.500 0.900
0.691 0.690
Wellbore (m)

600
Wellbore (m)

600
400 400 0.250 0.850
0.521 0.520
800 800
Wellbore (m)

Wellbore (m)

600 600 0.000 0.800


0.352 0.350

1000 1000
800 800
0.182 0.180

1200 1200
1000 1000

1400 1400
1200 1200
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Radial direction (m) Radial direction (m)
1400 1400
Figure 25: Normalized oil permeability after 24 hours production
RKO (left) and RKOE computed at maximum So (right), sim. FS4.
0.0 1.0 2.0 0.0 1.0 2.0
Radial direction (m) Radial direction (m)
Figure 22: Oil distribution around well after 200 hours mud
-1
exposure at 86 bar, 170 s (left) and after 24 hours back
production at 30 bar drawdown (right).
20 SPE 122241

So, 125 hrs, Sim. SF4b


0
So
0.860
200

0.696

400
0.532
Wellbore (m)

600
0.369

800
0.205

1000

1200

1400

0.0 0.5 1.0 1.5 2.0


Radial direction (m)
Figure 26: Sim. FS4b, with well penetration rate 10 m/hr.

Você também pode gostar