Anda di halaman 1dari 14

SPE 59537

Dynamic Model of Wormhole Formation Demonstrates Conditions for Effective Skin


Reduction During Carbonate Matrix Acidizing
C.N. Fredd, SPE, Schlumberger Oilfield Services

Copyright 2000, Society of Petroleum Engineers Inc.


This paper was prepared for presentation at the 2000 SPE Permian Basin Oil and Gas
Recovery Conference held in Midland, Texas, 2123 March 2000.
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, as
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
for commercial purposes 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 where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
The success of carbonate stimulation treatments is highly
dependent upon the formation of conductive flow channels, or
wormholes. Success requires wormhole formation during
matrix acidizing treatments and lack thereof during fracture
acidizing treatments. The structure of the wormhole channels,
which varies significantly with flow conditions and
acid/mineral properties, ultimately controls the effectiveness
of a stimulation treatment. This paper describes a dynamic
model of wormhole formation that accounts for wormhole
structure based on the Damkhler number. It addresses the
issue of scaling laboratory data to the field by including the
effects of fluid loss and wormhole competition. The model
predicts skin evolution based on basic input parameters and is
used to demonstrate matrix acidizing strategies for optimum
skin reduction with several acid systems.
The results demonstrate that, under typical treatment
conditions, conventional matrix treatments with straight HCl
cause face dissolution and provide little reduction of skin.
Under the same conditions, alternative fluids such as weak
acids and emulsified HCl create dominant wormhole
structures that penetrate deep into the formation. While these
fluids are more effective than straight HCl (especially at high
temperatures and low injection rates), additional injection
strategies can be used to increase the depth of penetration and,
in turn, improve skin reduction. Injection strategies such as
increasing the injection rate or decreasing the rate of
dissolution (by changing the fluid properties) promote deeper
penetration of wormholes. These injection strategies rely on
maintaining an optimum Damkhler number for effective
wormhole formation during matrix stimulation treatments. The

model also applies to the selection of fluid properties to


control wormhole formation during fracture acidizing
treatments.
Introduction
The transport and reaction of reactive fluids in carbonate
reservoirs results in the formation of highly conductive flow
channels, or wormholes. Wormholes significantly influence
the flow of reservoir fluids because their conductivity is
several orders of magnitude larger than that of the porous
medium. Therefore, the success of carbonate stimulation
treatments is highly dependent upon wormhole formation (i.e.,
wormhole formation during matrix acidizing treatments and
lack thereof during fracture acidizing treatments). Wormhole
formation is desirable during matrix acidizing treatments
because the wormholes are capable of bypassing nearwellbore damage. In contrast, wormhole formation increases
fluid leakoff during fracture acidizing and, consequently,
limits the depth of acid penetration. The structure of the
wormhole channel, which varies significantly with flow
conditions and fluid/mineral properties, ultimately controls the
effectiveness of stimulation treatments.
The importance of wormhole formation has led several
investigators to study the dissolution phenomenon. These
studies have provided a fundamental understanding of
wormhole formation and have resulted in models that predict
the rate of wormhole growth, type of wormhole structure, and
optimum
conditions
for
wormhole
formation.1-15
Unfortunately, scaling laboratory data to the field is not
straight forward, as noted by previous investigators.2-5
Wormhole formation in the field is complicated by the effects
of fluid loss through the walls of the wormhole in radial
geometry and by competition among wormholes for the
injected fluid. The effects of these processes have been
successfully included in capillary tube models.5,6 However,
these models were limited in their ability to simulate
wormhole formation because the effects of wormhole structure
were not taken into account.
Network models inherently include the effects of fluid
loss, wormhole competition, and wormhole structure because
pore-scale transport and reaction are accounted for in a
representative porous medium.1,2,7 Although the use of
network models would eliminate the need for scaling, the

C.N. FREDD

ability to simulate wormhole formation at the field scale is


limited by the excessive memory and computational time
required for such simulations. Thus, the most tractable
approach at this time is to scale theories of wormhole
formation to predict wormhole formation at the field scale.
This paper describes a dynamic model of wormhole
formation that is based on a capillary tube representation of
the porous medium. The model was introduced in a previous
publication9 and has been extended in this work. The effects of
wormhole structure are included through a dependence on the
Damkhler number and kinetic parameter.7,8 Wormhole
competition and fluid loss are included to scale the model to
the field scale. The model predicts wormhole formation and
skin evolution during matrix stimulation treatments under
typical field conditions. Limitations of conventional matrix
stimulation treatments are demonstrated and optimum
injection strategies to achieve effective skin reduction during
matrix stimulation treatments are discussed.
Wormhole Formation
Dissolution Structures. The structure of the dissolution
channels formed during transport and reaction in porous media
vary widely depending on parameters such as the injection rate
and fluid/mineral properties. The wide range of dissolution
structures is demonstrated in Fig. 1. The first four images are
neutron radiographs of dissolution structures formed during
the dissolution of calcite by HCl and EDTA.8,10 The fifth
image is a Woods metal casting of a dissolution structure
formed during the dissolution of dolomite by HCl.1 These
images illustrate the five main types of dissolution structures:
1.
2.
3.

Face dissolution
Conical wormholes
Dominant wormholes

4. Ramified wormholes
5. Uniform dissolution
The transition from dissolution structure 1 to 5 (from left to
right in Fig. 1) is commonly observed as the injection rate is
increased. At low injection rates, the reactant is consumed on
the inlet flow face of the core, resulting in face dissolution or
complete dissolution of the core starting from the inlet flow
face. This face dissolution (also referred to as compact
dissolution) consumes large volumes of reactant and provides
negligible depths of live acid penetration. At slightly higher
injection rates, the reactant can penetrate into the porous
matrix and enlarge flow channels. However, a significant
amount of reactant is consumed on the walls of the flow
channels, resulting in the formation of a conical-shaped
dissolution channel. At intermediate injection rates, reactant is
transported to the tip of the evolving flow channel, where
subsequent consumption propagates the channel and
eventually leads to the formation of a dominant wormhole. At
high injection rates, the dissolution channels become more
highly branched or ramified as fluid is forced into smaller
pores. At the extreme of very high injection rates, uniform

SPE 59537

dissolution is observed as the reactant is transported to most


pores in the medium.
The type of dissolution structure has a significant influence
on the volume of acid required to obtain a given depth of
wormhole penetration. This influence is demonstrated in Fig.
2, which shows the dependence of the number of pore
volumes to breakthrough, VBT, on the injection rate for the
dissolution of limestone by various stimulation fluids.7 All the
fluids exhibit an optimum injection rate at which the number
of pore volumes required to breakthrough is minimized and
dominant wormhole channels are formed. The number of pore
volumes to breakthrough increases to the left and right of the
minimum due to the formation of conical wormhole and
ramified wormholes, respectively. The optimum injection rate
corresponds to conditions at which a minimum volume of
fluid is required to achieve a given depth of wormhole
penetration. Hence, it represents the most effective conditions
for matrix stimulation. The existence of optimum injection
rates has been observed for a variety of fluid/mineral
systems.1-4,7,8,10-14
Dependence on the Damkhler Number. Numerous studies
have demonstrated the dependence of wormhole formation on
a variety of treatment variables and fluid/mineral properties.
The injection rate and effective diffusivity of the reactant have
a major influence on the dissolution phenomenon because of
the importance of mass transfer processes. The importance of
these parameters has led to models of wormhole formation
that depend on dimensionless terms such as the Damkhler
number1,7,8 and the Peclet number.2,3,5,11,13,15 Both types of
models successfully predict wormhole formation for at least a
few types of dissolution structures.16 However, only the model
by Fredd and Fogler7,8 accounts for all types of dissolution
structures and is valid for alternative stimulation fluids such as
organic acids and chelating agents. Therefore, the dependence
of wormhole formation on the Damkhler number, which
serves as the theoretical basis for this paper, is described in
this section.
The dependence of wormhole formation on the Damkhler
number was demonstrated for a wide range of fluid/mineral
systems including the dissolution of calcite by strong acids,
weak acids, and chelating agents, as well as the dissolution of
plaster of Paris by water.7,8 These systems are influenced by
various transport and reaction processes including the
transport of reactants to the mineral surface, reversible surface
reactions, and products transport away from the mineral
surface.17-21 The Damkhler number, NDa , is defined as the
ratio of the net rate of dissolution to the rate of transport by
convection and is given by
N Da =

dL
.....................................................................(1)
q

where q is the flow rate in the wormhole, d and L are the


diameter and length of the wormhole, respectively, and is
the overall dissolution rate constant. The overall dissolution

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

rate constant is a function of the various transport and reaction


processes and is given by
1+
=

1
K1

1
kr

1
K eq
+

.................................................(2)

1
K eq K3

where K1 and K3 are the mass transfer coefficients for


reactants and products, respectively, kr is the surface reaction
rate constant, Keq is the effective equilibrium constant for the
surface reaction, and is the stoichiometric ratio of reactants
consumed to products produced. The mass-transfer
coefficients were obtained from Levichs solution of the
convective diffusion equation for laminar flow in a cylindrical
tube.22 The average mass-transfer coefficient along the length
of a tube is given by
13

K mt =

1 .86 De2 3 4 q

d
L

..................................................(3)

where K mt is for either reactants or products (K1 or K3 ),


depending on the value of the diffusion coefficient. The
effective surface reaction rate constant and the effective
equilibrium constant were obtained from the independent
kinetic studies using a rotating disk.18-21
Physically, the Damkhler number defined in Eq. 1
provides a measure of the amount of reactant being consumed
on the walls of the wormhole, as opposed to being transported
to the tip of the wormhole where it can be consumed
efficiently. It dictates the type of wormhole structures that are
formed by systems with various degrees of transport and
reaction limitations. More importantly, there exists an
optimum Damkhler number at which dominant wormhole
channels are formed and a minimum volume of fluid is
required to obtain a given depth of wormhole penetration. This
optimum Damkhler number occurs at approximately 0.29 for
a wide range of fluid/mineral systems investigated in linear
coreflood experiments.
In addition to the dependence of wormhole formation on
the Damkhler number, Fredd and Fogler8 observed that the
dissolution phenomenon was also influenced by a
dimensionless kinetic parameter. The kinetic parameter, , is
defined as the ratio of the rate of surface reaction to the overall
rate of dissolution and is given by
= 1+

kr
kr
........................................................(4)
+
K1 Keq K3

They observed an optimum kinetic parameter at a value of


about 130 and presented a three-dimensional, cone-shaped
surface formed when the number of pore volumes to
breakthrough was plotted versus 1/NDa and (Fig. 3). The
surface demonstrates the optimum conditions for wormhole
formation and provides a means of predicting the dissolution
structure and number of pore volumes to breakthrough. Hence,
when used together, the Damkhler number and kinetic

parameter provide a complete description of wormhole


formation.
The Damkhler number theory does not independently
predict the rate of wormhole growth and, hence, needs to be
combined with a physical model of wormhole growth to
predict skin evolution. Moreover, the Damkhler number is
defined based on transport and reaction in an individual
wormhole, thus a method of determining the number of
wormholes is required to quantitatively apply the theory for
the optimum Damkhler number to the field scale.
Scaling Laboratory Data to the Field
Two main methods have been used to scale laboratory results
to the field. The simplest method involves scaling the
laboratory injection rate by the ratio of surface areas (i.e.,
maintaining the same superficial injection velocity).3,5,14 This
method inherently assumes that the same number of
wormholes will form per unit surface area in the field as in the
laboratory. In general, this assumption will over estimate the
number of wormholes that will form per unit surface area.
Therefore, the flow rate per wormhole and, consequently, the
rates of transport and reaction are not consistent with those
observed in the laboratory. Hence, this method does not
correctly scale the wormhole structure.
The second method involves scaling the injection rate by
the number of wormholes formed per unit surface area.14,23
This method maintains the same relative rates of transport and
reaction within the wormholes (i.e., maintains the same
Damkhler number) and, hence, maintains the same wormhole
structure. The importance of maintaining the same rates of
transport and reaction when scaling laboratory data is
demonstrated by results from laboratory studies. Different
wormhole structures were observed during linear coreflood
experiments when the cross-sectional area for flow was varied
and the superficial velocity was held constant.5 In contrast,
similar wormhole structures were observed when the crosssectional area was changed and the injection rate was held
constant.24 In both cases, a single wormhole channel was
observed to penetrate the length of the core. Therefore, the
rates of transport and reaction in the wormhole were kept
constant only in the second case. This comparison
demonstrates the importance of scaling based on the number
of wormholes to maintain the same wormhole structure.
Unfortunately, the number of wormholes per unit surface area,
or the wormhole density, is not easy to predict at the field
scale.
Recent studies have attempted to predict wormhole density
using capillary tube models.23,25 Huang et al.23 solved the
steady state pressure diffusivity equation for fluid loss from a
static capillary and assumed wormholes could not grow until
the pressure gradient reached an initiation pressure gradient.
Based on this assumption, they predicted wormhole densities
for specific depths of wormhole penetration. Gdanski25
predicted wormhole density based on questionable arguments
of symmetry and simplifying assumptions that strongly
influenced the model output. Unfortunately, many of the
model predictions contradicted well accepted bodies of work

C.N. FREDD

in the literature. Neither model accounts for the effects of


wormhole structure on wormhole density. Therefore, although
these models address a major gap in our understanding of
wormhole formation, additional studies are required to fully
understand the effects of stimulation variables on wormhole
density.
Dynamic Model of Wormhole Formation
The dynamic model of wormhole formation described in this
section was introduced in a previous publication9 and is
extended in this work. The model is based on a capillary tube
representation of the dominant dissolution channels. Fluid loss
from the walls of the dissolution channels is included and the
effects of transport and reaction on channel evolution are
determined based on the Damkhler number and kinetic
parameter. Thus, all three transport and reaction processes are
taken into account. The use of these parameters allows for
determination of the type of dissolution structure that is
formed. This information is then used to calculate the axial
and radial wormhole growth rates to account for the effects of
wormhole structure. This physical model of wormhole growth
is combined with predictions of wormhole density and
calculations of wormhole competition to scale the model to the
field.
The dependence of wormhole formation on the Damkhler
number and the kinetic parameter was observed in laboratory
experiments conducted in linear cores.7,8 In those experiments,
only one wormhole was typically observed due to the limited
cross-sectional area available for flow. In addition, essentially
all of the flow was through that single wormhole because the
conductivity of the wormhole was several orders of magnitude
higher than that of the porous medium and fluid loss through
the walls of the wormhole was insignificant in the linear
geometry. Therefore, the experiments demonstrated the
optimum Damkhler number and optimum kinetic parameter
within a single wormhole with negligible fluid loss. To predict
wormhole formation in the field, the effects of fluid loss and
wormhole competition must be taken into account.
Fluid loss and wormhole competition both affect the flow
rate per wormhole and, consequently, affect the rates of
transport and reaction. Because the Damkhler number and
kinetic parameter are defined based on transport and reaction
in a single wormhole, the effects of these processes can be
easily accounted for as long as the flow rate per wormhole can
be determined. The scaling method based on wormhole
density is directly applicable to solving this problem.
Knowledge of the total injection rate and wormhole density,
along with the fluid loss rate per wormhole, provides the flow
rate per wormhole. Therefore, it is reasonable to assume that
the Damkhler number and kinetic parameter will describe
wormhole formation at the field scale, provided the effects of
fluid loss and wormhole competition are taken into account.
In describing the dynamic model of wormhole formation,
the following sections extend the theory of the Damkhler
number7,8 to include the effects of fluid loss from the walls of
the wormhole, introduce a method to account for wormhole
competition, and describe the modeling strategy.

SPE 59537

The Damkhler Number and Fluid Loss. Fluid loss through


the walls of a wormhole has three main effects on transport
and reaction within the wormhole. These effects include
decreasing the flow rate along the length of the wormhole,
affecting the rates of mass transfer within the wormhole (due
to the combined influence of convection and diffusion normal
to the walls of the wormhole), and increasing the amount of
reactant lost through the walls of the main wormhole channel
(due to reactant leakoff). Each of these effects reduces the
amount of reactant that is transported to the tip of the
wormhole and, consequently, reduces the rate of wormhole
propagation.
To simulate the effects of fluid loss on the concentration of
reactants transported along the wormhole, fluid loss was added
to the model for the dissolution of a representative cylindrical
tube developed by Fredd and Fogler.7,8 The cylindrical tube
represents the dominant flow channels within the porous
medium (i.e., the wormholes). Convection (in the axial and
radial directions) and reaction in the tube were included in a
reactant mass balance. To obtain an analytical solution, the
concentration of reactants leaking off into the formation was
assumed to be equal to the concentration at the solid-liquid
interface. This assumption is valid for three cases: zero fluid
loss velocity, surface reaction limited dissolution, and masstransfer limited dissolution. This assumption is commonly
used when modeling fracture acidizing and has been described
in more detail in the literature.26 The resulting expression for
the concentration profile along the length of the flow channel,
which is similar to that reported by Fredd and Fogler, is given
by
C
( N + )
= (1 B ) e Da
+ B ...............................................(5)
Co
where,

v
1 l

K1
N Da

B=
.................................................(6)
( N Da + ) 1 + K eq

The exponent in Eq. 5 includes the Damkhler number as


defined in Eq. 1 and an additional dimensionless fluid loss
term () that is given by
=

dLvl
q


1
..........................................................(7)

K1

When applying Eqs. 1 through 7, all mass transfer coefficients


are corrected to account for the combined influence of
convection and diffusion normal to the walls of the wormhole.
For rapid mass transfer due to both diffusion and convection,
the corrected mass-transfer coefficient (K) is given by27
v
K = (v K l )
...............................................................(8)
l
mt 1
e

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

where vl is the fluid loss velocity and Kmt is the mass-transfer


coefficient used in the absence of fluid loss (i.e., Eq. 3). The
fluid loss velocity is negative for flow in the direction of
diffusion (i.e., for the transport of reactants to the surface). At
typical reservoir conditions, the corrected and uncorrected
mass-transfer coefficients differ by less than an order of
magnitude.
The obvious importance of the fluid loss velocity is
demonstrated by the fluid loss term (Eq. 7). As the fluid loss
velocity decreases, the influence of the fluid loss term
decreases and the Damkhler number eventually dominates
the dissolution. Such is the case in linear coreflood
experiments in the laboratory. Not so obvious is the
dependence of the fluid loss term on the degree of masstransfer limitations. When the dissolution is limited by the
transport of reactants to the surface ( = K1 ), the fluid loss
term is zero because the concentration of reactants at the solidliquid interface (and, consequently, the concentration of
reactants leaking off into the formation) is negligible. Thus,
for a reactants transport limited system like HCl/limestone, the
dissolution is dependent on transport and reaction within the
wormhole and, hence, the Damkhler number dictates the
reactant consumption. [For example, NDa /(NDa + ) = 0.999 for
the HCl/limestone system at 200F, even with a high fluid
velocity.] On the other hand, for fluid/mineral systems that are
influenced by the kinetics of the surface reaction (e.g.,
HCl/dolomite and EDTA/limestone at ambient temperature),
the fluid loss term and the Damkhler number are both
significant. Thus, fluid loss does not influence the dependence
of wormhole formation on the Damkhler number when the
fluid loss term is low (i.e., in damaged formations with low
permeability, when fluid loss additives are present, or when
the wormholes are relatively short) and when the dissolution is
limited by the transport of reactants to the surface. Note that
these conditions are consistent with the assumptions used in
deriving the analytical solution to this problem.
In this model, Eqs. 5 through 8 are used to account for the
effects of fluid loss on reactant consumption based on the
Damkhler number and fluid loss term. The Damkhler
number and kinetic parameter are used to predict the
dissolution structure. The impact of fluid loss on the
dissolution structure is not known, but it is possible that the
dissolution structure depends on NDa + and an optimum NDa
+ may exist. Although not addressed in this work, this
possibility warrants future investigations.
Wormhole Competition. During field treatments, many
wormholes (or other types of dissolution structures) will form
and compete for the injected fluids. The longer wormholes,
which typically have larger diameters, will accept more fluid
than the shorter wormholes. As a result, the longer wormholes
will propagate more rapidly, while the shorter wormholes will
eventually stop growing as they receive an insufficient amount
of reactant.1,6 The number of wormholes that are capable of
penetrating to a given depth depends on the length of the
competing wormholes and their separation distance. For

example, results from simulations in linear geometry have


shown that wormhole interactions significantly reduce the
flow rate in a wormhole when the distance between
neighboring wormholes is less than the wormhole length.1,5 In
general, the number of relatively long wormholes will
decrease as the depth of penetration increases. Thus,
wormhole competition and wormhole density will change
throughout the stimulation treatment.
To simulate the effects of wormhole competition, the
fraction of the fluid entering the dominant dissolution
channels (fd ) was determined. Assuming radial flow and the
same dimensions for all the dominant dissolution channels, fd
is given by
1
Qt
1
=
= 1+
fd
nwh qwh
nwh q wh

n wh( short)

qwh (short) (i ) ...............(9)


i =1

where n wh is the number of dominant dissolution channels (i.e.,


those controlling the skin evolution). The subscript short
indicates shorter dissolution channels that are not influencing
the skin value, but are competing for injected fluids.
Substituting Darcys law for radial flow for the inlet flow rates
per dissolution channel and assuming the pressure profile
around a short wormhole is not influenced significantly by a
neighboring dominant wormhole, Eq. 9 reduces to the
following expression:
rf
ln
r
1
= 1 + wh
fd
nwh

n
wh ( short)

i =1

1
...........................(10)
rf

ln

rwh (i)

where rwh = rw + L is the radius of dominant dissolution


channels, rw is the radius of the wellbore (6 in.), and rf is the
radius of the reservoir (1,000 ft). If the length of a short
dissolution channel is less than 10% of the length of the
dominant dissolution channels, then the flow rate in the short
dissolution channel is assumed to be zero. This assumption is
intended to compensate for the lack of influence of pressure
profiles around neighboring wormholes.
The number of wormholes or wormhole density is required
to account for wormhole competition. Although no study has
reported the wormhole density as a function of injection
conditions and depth of penetration, this functionality can be
estimated from information reported in the literature. The
results of network model simulations indicate that the number
of wormholes should scale roughly with the ratio of the inlet
area to the depth of penetration.1 A similar observation was
made from a capillary tube model.5 The scaling rule observed
by Hoefner and Fogler1 was used to account for the effects of
the depth of penetration in this study. Although the scaling
rule was observed for dissolution structures consistent with
dominant wormholes, this scaling rule was assumed to be
valid for all types of dissolution structures. It should be noted
that this is not the case for uniform dissolution. The

C.N. FREDD

assumption does not influence results when predicting


optimum injection strategies for efficient wormhole formation.
In addition to being influenced by the depth of penetration,
the number of dominant dissolution channels will depend on
the Damkhler number within each wormhole. At the
extremes of high and low Damkhler numbers, face
dissolution and uniform dissolution will result in a single
dissolution structure and an infinite number of dissolution
channels, respectively. Between these extremes, a gradual
increase in the wormhole density is expected as Da is
decreased. This influence of the Damkhler number (or
wormhole structure) on wormhole density has not been
considered in previous studies. Therefore, estimates of the
dependence of wormhole density on the Damkhler number
and kinetic parameter were made based on laboratory
observations for straight and emulsified HCl. An assumed
wormhole density is shown in Fig. 4 as a function of NDa and
. The contour represents the number of wormholes that are
expected to exist on a 12 in.2 surface when the depth of
penetration is about 4 in. The functionality was assumed to be
independent of the depth of wormhole penetration and,
therefore, was scaled with changes in depth of penetration
based on the scaling rule observed by Hoefner and Fogler.
This method of calculating the number of wormholes is
applicable for an open-hole completion and represents the
upper limit for the wormhole density. The calculations can be
modified to include the effects of perforations on wormhole
density by limiting the maximum number of wormholes to the
number of perforations. Huang et al.14 introduced a similar
approach, but assumed a constant number of wormholes (i.e.,
one wormhole per perforation).
Description of Wormhole Model. The dynamic model of
wormhole formation simulates the growth and competition of
dissolution channels. The model is based on three main
assumptions: (1) the dependence of wormhole formation on
the Damkhler number and kinetic parameter is not affected
by fluid loss or wormhole competition, (2) the rate of fluid
loss is consistent with that predicted for a single wormhole,
and (3) the wormhole density is consistent with that estimated
from laboratory data and follows the scaling rule observed
from network model simulations. Because these assumptions
must be verified experimentally, the results presented in this
paper are considered qualitative trends.
The model simulates wormhole formation under two
conditions: constant injection rate and constant Damkhler
number. The constant Damkhler number condition provides a
means of predicting optimum injection strategies for matrix
stimulation treatments. Because this is a novel approach, the
model will be described in detail for the constant Damkhler
number condition. The model differs only by the solution
strategy for the two conditions.
Constant Damkhler Number. The predictions of the
optimum injection conditions for matrix treatments are based
on determining the conditions required to maintain an

SPE 59537

optimum Damkhler number of 0.29. Because this Damkhler


number corresponds to the formation of dominant wormhole
structures, the wormhole density function introduced in the
previous section does have a significant influence on the
results. Under these conditions, the wormhole density is
influenced mainly by the scaling rule for the effects of depth
of penetration. It was assumed that it is necessary to maintain
the optimum Damkhler number in only the dominant
dissolution channels (i.e., the longer wormholes that carry
most of the injected fluid, as shown in Fig. 5). Therefore,
transport and reaction in the shorter dissolution channels was
not taken into account. Once a dissolution channel became
relatively short or non-dominant, its geometry remained
constant and only competition for fluid with that wormhole
was included for the remainder of the simulation. The
calculations used to estimate the relative flow rate in the
dominant dissolution channels were described in the previous
section.
The model focuses on transport and reaction in the
dominant dissolution structures. Therefore, the calculations
depend on the number of dominant wormholes that can
penetrate to a depth L. Because the number of wormholes,
diameter of the wormholes, growth rate, and fluid loss rate
change as the depth of penetration increases, the wormhole
properties are evaluated for discrete increases in the depth of
penetration. The solution algorithm consists of the following
seven steps:
Step 1. The fluid/mineral system and temperature are
specified and the relevant parameters are evaluated (i.e.,
diffusion coefficients, surface reaction rates, and effective
equilibrium constants). The initial dissolution channel length
(0.02 in.) and diameter (20 m) are specified. (Simulations
were not sensitive to perturbations of the initial channel
geometry.)
Step 2. The optimum injection rate in a single wormhole
of length L is calculated from Eq. 1. This calculation requires
an iterative solution involving calculation of the mass transfer
coefficients and the rate of fluid loss. The rate of fluid loss
was estimated from predictions for fluid loss in a single
wormhole.15
Step 3. The wormhole growth rate is calculated by
assuming the reactant transported to the tip of the wormhole
will propagate the wormhole channel by dissolving the tip
over the same cross-sectional area as the existing wormhole.
The growth rate is given by
dL
MWA
= rAt
X t BT ...................................................(11)
dt
A
where rAt is the rate of reactant consumption at the tip, MWA
and A are the molecular weight and density of the reactant,
respectively, and Xt is the volumetric dissolving power
(volume of mineral dissolved by a given volume of reactant)
at the tip. BT is a structure efficiency parameter that accounts

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

for the effects of dissolution structure. This parameter will be


discussed in more detail.
The rate of reactant consumption at the tip is given by7

Co
rA = C
.................................................... (12)
1 + Keq

where the overall dissolution rate constant was calculated at


the tip of the wormhole by replacing the mass-transfer
coefficients in Eq. 2 by the superficial velocity at the tip. With
the exception of accounting for the effects of wormhole
structure, Eq. 11 is similar to the growth rate expression used
by Hung et al.6 In addition, it accounts for both transport and
reaction processes at the tip and, therefore, allows for
unconsumed reactant to be transported beyond the tip of the
dissolution channel (which may occur when the dissolution is
influenced by the kinetics of the surface reaction).
The effects of dissolution structure on the wormhole
growth rate were included by introducing a structure
efficiency parameter,
BT

= VBT ( opt)
V
BT

if N DA > 0. 29 and > 130


all other conditions

............(13)

where VBT(opt) is the minimum number of pore volume required


to breakthrough at the optimum stimulation conditions (i.e., at
the optimum NDa and ). Therefore, BT scales the rate of
propagation of a particular dissolution structure relative to the
rate of propagation of an optimum dissolution structure. When
NDa > 0.29 and > 130, the dissolution structure resembles a
capillary tube (i.e., exhibits minimum branching) and no
correction for structure is required. For all other conditions,
the dissolution structure deviates from a capillary tube because
of branching and the correction for structure was applied. VBT
was determined by the fitting the three-dimensional surface in
Fig. 3 as a function of NDa and . The number of pore volumes
to breakthrough is inversely related to the average rate of
wormhole propagation for a particular depth of penetration.
Although this relationship will change with depth of
penetration, the number of pore volumes to breakthrough was
used as a first approximation to include the effects of
dissolution structure.
Step 4. The change in time associated with the step change
in wormhole length is calculated from the wormhole growth
rate (dL/dt)
dt =

L
.......................................................................(14)
dL

dt

This step is necessary because the solution strategy relies on


specifying the step size based on discrete increments in the
wormhole length.

Step 5. Once the time step size is determined, the volume


of fluid injected into the wormhole is calculated, and the
change in the wormhole diameter is determined from the
amount of reactant consumed on the walls of the main
wormhole channel (c.f., Eq. 5). Reactant leakoff will lead to
additional dissolution of the rock matrix surrounding the
wormhole, but in many cases the dissolution will occur in side
branches, not the main flow channel. To account for this
effect, reactant consumption was scaled by the structure
efficiency parameter defined in Eq. 13. Thus, the structure
efficiency parameter was used to account for the effects of
dissolution structure on both the axial and radial rates of
wormhole growth.
Step 6. The injection rate determined in step 2 represents
an average injection rate along the length of the dominant
wormhole that is required to maintain the optimum Damkhler
number in that wormhole. To achieve that average injection
rate, the injection rate into the wormhole must be larger to
compensate for fluid loss from the walls of the wormhole.
This required injection rate per dominant dissolution channel,
q wh , was approximated as
q wh = q + ql .................................................................... (15)
where q is the average flow rate per dominant dissolution
channel (determined based on NDa using Eq. 1) and q l is the
rate of fluid loss. The total injection rate (Qt ) is then calculated
from
Qt = nwh q wh f d ..............................................................(16)
where fd is given by Eq. 10.
Step 7. The length of the dominant dissolution channels is
incremented and steps 2 through 7 are repeated until the
desired depth of wormhole penetration or volume of injected
fluid is achieved.
The model predicts the optimum injection rate (for NDa =
0.29) as a function of depth of wormhole penetration or
volume injected. Similarly, the model can predict optimum
fluid properties to maintain the optimum Damkhler number
and a constant injection rate. Using the same set of equations
and a slightly different solution strategy, the model also
predicts changes in the Damkhler number and dissolution
structure associated with injecting at a constant rate. In all
cases, the wormhole length is predicted as a function of the
volume of fluid injected. Based on this information, the skin
evolution is calculated using the three-zone model described
by Frick et al.15
Results and Discussion
Before presenting optimum injection strategies for field
treatments, it is first necessary to understand how the various
dissolution structures, and hence the Damkhler number,
influence the effectiveness of matrix stimulation treatments.
Typical skin evolution curves are shown in Fig. 6 for various

C.N. FREDD

dissolution structures created at constant Damkhler number.


The skin evolution varies significantly with dissolution
structure due to changes in the depth of penetration (Fig. 7).
When face dissolution occurs, the skin reduction is
insignificant as acid penetrates only a fraction of an inch into
the formation. Conical wormholes result in a slight decrease in
skin and about 2 inches of penetration. Near the optimum
Damkhler number, wormhole formation results in effective
stimulation as evidenced by the negative skin. Dominant
wormholes were observed to penetrate over 12 inches into the
formation. As the Damkhler number was decreased and the
dissolution structure changed to ramified wormholes and
uniform dissolution, the skin reduction became less effective
and the depth of penetration decreased. This trend is consistent
with experimental and theoretical results reported by several
investigators.1-15
The formation of dominant wormhole channels represents
the most effective mode of stimulation for a given volume of
fluid injected. This effectiveness is due to the dominant
wormhole channels providing the greatest depth of
penetration, as shown schematically in Fig. 8. Obviously the
other types of dissolution structures are capable of stimulation
beyond the damaged zone if a sufficient volume of fluid is
injected. The trade off is the cost of the additional fluid
injected and possible loss of integrity of the near-wellbore
matrix due to excessive dissolution.
The wormhole density varied significantly with changes in
wormhole structure and depth of penetration (Fig. 9). Only
one dissolution structure was observed per square foot of
wellbore surface area for face dissolution, whereas about
10/ft 2 were observed for uniform dissolution. Once again, it
should be noted that the wormhole scaling methods are not
strictly correct for uniform dissolution. Interestingly, the
wormhole density for uniform dissolution at a penetration
depth of 6 in. was consistent with that predicted by Huang et
al.23 under similar conditions.
The wormhole densities shown in Fig. 9 are the result of
the scaling rules and density function used in the model. To
demonstrate the sensitivity of the simulations to these inputs,
the effects of changes in wormhole density on skin reduction
for a constant injection rate are shown in Fig. 10. A significant
decrease in the rate of skin reduction was observed when the
number of wormholes was increased by a factor of ten. In
contrast, a slight increase in the rate of skin reduction was
observed when the number of wormholes was decreased by a
factor of ten. Thus, the number of wormholes has a significant
effect on predictions of skin evolution.
The trends observed in Fig. 10, as well as the average
values of the Damkhler number for the three cases, are
consistent with a transition from dominant wormholes to
conical wormholes as the number of wormholes was
increased. The skin reduction became more effective as the
number of wormholes decreased and, correspondingly, the
flow rate per wormhole increased. Simulations performed with
alternative fluids such as formic acid (HFc), acetic acid (HAc),
and emulsified HCl resulted in more effective skin reduction
than HCl under the same conditions (Fig. 11). Although these

SPE 59537

fluids were more effective, these simulations were not run at


the optimum stimulation conditions.
Optimum Injection Strategies. Optimum injection strategies
for field treatments were predicted for an optimum Damkhler
number of 0.29. This optimum condition can be maintained by
increasing the injection rate as the depth of penetration
increases. This approach is demonstrated in Fig. 12, which
shows the normalized optimum injection rate required to
maintain the optimum Damkhler number as a function of the
depth of penetration. (The optimum injection rate is
normalized by the optimum injection rate for HCl at 200F
with zero fluid loss.) The curve represents injection rates at
which efficient wormhole formation will occur. At injection
rates above and below the optimum curve, ramified
wormholes and face dissolution will form, respectively. The
curves reveal that the injection rate must be increased
significantly as the depth of penetration increases. This
increase in injection rate is necessary to offset the effects of
fluid loss from the wormhole channels. When the fluid loss
velocity is low, the rate at which the injection rate must be
increased is less significant. The need to increase the injection
rate to maintain efficient wormhole formation is consistent
with investigators reporting a higher optimum injection rate in
radial experiments than in linear experiments.13,15
The effect of fluid type on the injection strategy in
limestone formations at 200F is shown in Fig. 13. The shaded
box represents typical injection rates used in conventional
matrix stimulation treatments.28 The results show that HCl is
unable to achieve significant penetration without requiring
excessive injection rates that would fracture the formation. In
contrast, alternative fluids such as weak acids and emulsified
HCl can stimulate to increasingly deeper depths without
exceeding the same maximum injection rate. Under these
conditions, emulsified HCl would be the most effective
stimulation fluid. Results for EDTA are similar to those of
emulsified HCl. The data also demonstrate that weak acids
and emulsified HCl are more effective than aqueous HCl when
the treatments are limited to low injection rates. This
effectiveness at low injection rates is consistent with results
reported by previous investigators.7,8,10,21,29,30
Fig. 13 reveals an alternative injection strategy, as
indicated by the dashed arrow. This strategy involves
maintaining a constant injection rate and gradually changing
the reactant type (thereby changing the overall rate of
dissolution). For example, a more effective stimulation could
be achieved by injecting a HCl/HAc blend that is gradually
changed from HCl to HAc as the depth of penetration
increases.
The optimum injection rate is a strong function of
temperature, as shown in Fig. 14. To obtain a particular depth
of penetration, the optimum injection rate must be increased as
the temperature increases. This trend is consistent with
experimental results with HCl where increasing the
temperature resulted in an increase in the optimum injection
rate.8,12 A much more significant dependence of the optimum
injection rate on temperature was predicted by Huang et al.14

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

due to the emphasis they placed on the rate of the surface


reaction rather than the rate of mass transfer. Fig. 14
demonstrates that optimal stimulation with HCl is limited to
low temperatures if excessive injection rates and,
consequently, fracturing the formation are to be avoided. The
figure also reveals that weak acids and emulsified HCl are
more effective than aqueous HCl when stimulating high
temperature limestone formations.
Conclusions
1. Optimum injection strategies and limitations of
conventional matrix stimulation treatments were
demonstrated using a dynamic model of wormhole
formation. The model includes the effects of fluid loss
and wormhole competition in an extension of the theory
for the optimum Damkhler number.
2.

3.

4.

5.

To maintain efficient wormhole formation, the


Damkhler number is maintained at its optimum value by
either increasing the injection rate or decreasing the
overall dissolution rate (by changing the fluid properties)
as the depth of penetration increases.
Under typical matrix stimulation conditions, treatments
with HCl are optimal only in low temperature formations
or at excessively high injection rates. Fluids, such as weak
acids and emulsified HCl, provide effective alternatives to
HCl at high temperatures and low injection rates.
The results discussed in this paper provide qualitative
predictions of optimum injection strategies that are in
agreement with experimental observations.
The simulations revealed the importance of fluid loss and
wormhole competition on skin evolution. The significant
dependence of model predictions on these processes
demonstrates the need for a rigorous investigation of their
effects on wormhole formation.

Nomenclature
B = constant
C = concentration, mol/L3 , mol/L
d = wormhole diameter, L, cm
De = effective diffusion coefficient, L2 /t, cm2 /s
fd = fraction of total flow entering the dominant
dissolution channels
kr = effective surface reaction rate constant, L/t, cm/s
K = mass-transfer coefficient, L/t, cm/s
Keq = effective equilibrium constant
L = wormhole length, L, cm
MWA = molecular weight of reactant, M/mol, g/mol
n wh = number of dominant dissolution channels
NDa = Damkhler number
q = flow rate per wormhole, L3 /t, cc/min
q l = fluid loss rate, L3 /t, cc/min
q wh = injection rate in a single wormhole with fluid loss,
L3 /t, cc/min
Qt = total injection rate, L3 /t, cc/min
rA = rate of reactant consumption, mol/L2 /t, mol/cm2 /s
rf = radius of reservoir, L, ft

rw = radius of wellbore, L, in.


rwh = radius of wormhole penetration, L, in.
t = time
VBT = number of pore volumes to breakthrough
vl = fluid loss velocity, L/t, cm/s
X = volumetric dissolving power
= overall dissolution rate constant, L/t, cm/s
BT = structure efficiency parameter
= kinetic parameter
= stoichiometric ratio of reactants to products
A = density of reactant, M/L3 , g/cm3
wh = wormhole density, 1/L2 , #/cm2
= fluid loss term
Subscripts
o = initial
1 = reactants
3 = products
mt = mass transfer
opt = optimum injection conditions
short = short or non-dominant dissolution channels
t = tip of wormhole

Acknowledgments
The author thanks Matt Miller for useful discussions and
Schlumberger for giving permission to publish the results of
this work.

References
1. Hoefner, M.L. and H.S. Fogler Pore Evolution and Channel
Formation During Flow and Reaction in Porous Media, AIChE
J. 34 (1), 45 (1988).
2. Daccord G., E. Touboul, and R. Lenormand Carbonate
Acidizing: Toward a Quantitative Model of the Wormholing
Phenomena, SPE Production Eng., 63 (February 1989).
3. Frick, T.P., B. Mostofizadeh, and M.J. Economides Analysis of
Radial Core Experiments for Hydrochloric Acid Interaction
With Limestones, paper SPE 27402 presented at the SPE
International Symposium on Formation Damage Control,
Lafayette, LA, February 7-10 (1994).
4. Bazin, B., C. Roque, and M. Bouteca A Laboratory Evaluation
of Acid Propagation in Relation to Acid Fracturing: Results and
Interpretation, paper SPE 30085 presented at the European
Formation Damage Conference, The Hague, The Netherlands,
May 15-16 (1995).
5. Buijse, M.A. Understanding Wormholing Mechanisms Can
Improve Acid Treatments in Carbonate Formations, paper SPE
38166 presented at the European Formation Damage
Conference, The Hague, The Netherlands, June 2-3 (1997).
6. Hung K.M., Hill A.D., and Sepehrnoorl K.; A Mechanistic
Model of Wormhole Growth in Carbonate Matrix Acidizing and
Acid Fracturing, J. Pet. Tech., 59, (January 1989).
7. Fredd, C.N. and H.S. Fogler, Influence of Transport and
Reaction on Wormhole Formation in Porous Media, AIChE J.,
44 (9) 1933-1949 (Sept. 1998).

10

C.N. FREDD

8. Fredd, C.N., and Fogler, H.S.: "Optimum Conditions for


Wormhole Formation in Carbonate Porous Media: Influence of
Transport and Reaction", SPE J., 4 (3), (Sept. 1999).
9. Fredd, C.N.: "Advances in Understanding and Predicting
Wormhole Formation," Reservoir Stimulation, 3rd Edition, K.
Nolte and M. Economides (eds.), John Wiley and Sons (1999).
10. Fredd, C.N. and H.S. Fogler Alternative Stimulation Fluids and
Their Impact on Carbonate Acidizing, SPE J. 13 (1), 34
(March 1998).
11. Daccord, G., Lenormand, R., and Lietard, O., Chemical
Dissolution of a Porous Medium By A Reactive Fluid - I.
Model for the Wormholing Phenomenon, Chem. Eng. Sci.
48, No. 1, 169-178 (1993).
12. Wang, Y., A.D. Hill, and R.S. Schechter The Optimum
Injection Rate for Matrix Acidizing of Carbonate Formations,
paper SPE 26578 presented at the Annual Technical Conference
and Exhibition, Houston, TX, October 3-6 (1993).
13. Mostofizadeh, B. and M.J. Economides Optimum Injection
Rate From Radial Acidizing Experiments, paper SPE 28547
presented at the SPE 69th Annual Technical Conference and
Exhibition, New Orleans, LA, September 25-28 (1994).
14. Huang, T., A.D. Hill, and R.S. Schechter Reaction Rate and
Fluid Loss: The Keys to Wormhole Initiation and Propagation
in Carbonate Acidizing, paper SPE 37312 presented at the
International Symposium on Oilfield Chemistry, Houston, TX,
February 18-21 (1997).
15. Frick, T.P., M. Krmayr, and M.J. Economides Modeling of
Fractal Patterns in Matrix Acidizing and Their Impact on Well
Performance, SPE Prod. and Facilities, 61-68 (February 1994).
16. Fredd, C.N. and Miller, M.J.: Validation of Carbonate Matrix
Stimulation Models, SPE 58713 to be presented at the SPE
International Symposium on Formation Damage Control,
Lafayette, LA, February 23-24, 2000.
17. de Rozieres, J., Chang, F.F., and Sullivan, R.B., Measuring
Diffusion Coefficients in Acid Fracturing Fluids and Their
Application to Gelled and Emulsified Acids, paper SPE 28552
presented at the SPE Annual Technical Conference and
Exhibition, New Orleans, LA, September 25-28 (1994).
18. Lund, K., H.S. Fogler, C.C. McCune, and J.W. Ault
Acidization - II. The Dissolution of Calcite In Hydrochloric
Acid, Chem. Eng. Sci. 30, 825 (1975).
19. Fredd, C.N. and H.S. Fogler The Influence of Chelating Agents
on the Kinetics of Calcite Dissolution, J. Colloid Interface Sci.,
204 (1), 187-197 (August 1998).
20. Fredd, C.N. and H.S. Fogler "The Kinetics of Calcite
Dissolution in Acetic Acid Solutions", Chem. Eng. Sci., 53 (22),
3863-3874 (October 1998).
21. Takulpakdee, S., Flo w and Reaction of Weak Acids in
Carbonate Porous Media, M.S. Thesis, Chulalongkorn
University, Bangkok, Thailand, (1998).
22. Levich, V.G., Physicochemical Hydrodynamics, Prentice-Hall,
Englewood Cliffs, NJ (1962).
23. Huang, T., Zhu, D., and Hill, A. D.: Prediction of Wormhole
Population Density in Carbonate Matrix Acidizing, SPE 54723,
presented at the European Formation Damage Conference, The
Hague (May 31-June 1, 1999).
24. Fredd, C.N., Unpublished Data, 1998.
25. Gdanski, R. D.:
A Fundamentally New Model of Acid
Wormholing in Carbonates, SPE 54719, presented at the

26.
27.
28.
29.

30.

SPE 59537

European Formation Damage Conference, The Hague (May 31June 1, 1999).


Settari A., Modeling of Acid-Fracturing Treatments, SPE
Production & Facilities, 30, (February 1993).
Cussler, E.L., Diffusion: Mass Transfer in Fluid Systems,
Cambridge University Press, New York (1984).
Paccaloni, G., A New, Effective Matrix Stimulation Diversion
Technique, SPE Prod. and Facilities, 151-156 (August 1995).
Hoefner, M.L. and Fogler, H.S., Effective Matrix Acidizing in
Carbonates Using Microemulsions, Chem. Eng. Prog. 40-44,
(May 1985).
Bazin, B.: "Experimental Investigation of Some Properties of
Emulsified Acid Systems for Stimulation of Carbonate
Formations", paper SPE 53237 presented at the 1999 Middle
East Oil Show, Bahrain, February 20-23.

SI Metric Conversion Factors


bbl 1.589 873
E-01 = m3
ft 3.048*
E-01 = m
ft 2 9.290 304* E-02 = m2
o o
F( F-32)/1.8
= oC
gal 3.785 412
E-03 = m3
in. 2.54*
E+00 = cm
md 9.868 233
E-04 = m2
psi 6.894 757 E+00 = kPa
*Conversion factor is exact.

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

Face
Dissolution10

Conical
Wormhole8

Dominant
Wormhole10

Ramified
Wormhole10

11

Uniform
Dissolution1

0.25 M DTPA, pH=4.3


0.25 M EDTA, pH=13
0.25 M EDTA, pH=4
0.25 M CDTA, pH=4.4
0.5 M HAc
0.5 M HCl

100

P ore volumes to breakthrough

Pore volumes to breakthrough

Fig. 1 Wide range of dissolution structures observed in linear coreflood experiments.

10

1
10

10

-2

250

100

100

10

10

1
0.5

0.5

length = 4 in.
diameter = 1.5 in.
-3

250

0.3

10

-1

10

10

10

1
10

Injection rate [cm /min]


Fig. 2 Dependence of the number of pore volumes to
breakthrough on the injection rate for various stimulation fluids.7

1/N Da

100

1000
1000 0

100

10

Reaction limited

Transport limited

Fig. 3 Optimum injection conditions for wormhole formation.8

C.N. FREDD

80

80

70

70

SPE 59537

60

60

40

50

30

40

50

20

30

10

20

0
0
8
3.4

S9

1.6

S7

S5

S3

S1

Uniform dissolution

Conical wormholes

2
Ramified wormholes

1
0

Dominant wormholes
-2

0
12 -1.2

Face dissolution

-1

10

4
Ln( )

Skin [dimensionless]

Number of wormholes

12

-3

5.7

Ln(1/NDa)

Fig. 4 Dependence of the number of wormholes on the


Damkhler number and kinetic parameter (surface area = 12 in.2
and depth of penetration = 4 in.).

10

Volume injected [bbl]

Fig. 6 Effects of dissolution structure on skin evolution at


constant Damkhler number.

Convection to tip of
wormhole
14
Dominant wormholes
12
Depth of penetration [in]

Fluid loss through


walls of wormhole

10
8
6

Ramified wormholes

4
Conical wormholes
2
Face dissolution / Uniform dissolution
0

Competition for
injected fluid
Fig. 5 Idealized representation of fluid loss from the walls of
wormholes and competition among wormholes for injected fluid.

10

Volume injected [bbl]

Fig. 7 Effects of dissolution structure on the depth of


penetration.

SPE 59537

DYNAMIC MODEL OF WORMHOLE FORMATION FOR CARBONATE MATRIX ACIDIZING

13

Damaged zone

Face
dissolution

Conical
wormhole
6

Skin [dimensionless]

5
Dominant
wormhole

Ramified
wormholes

4
3

nwh *10

2
nwh

1
nwh /10

0
-1
-2
-3

Uniform
dissolution

10

Volume injected [bbl]

Fig. 8 Schematic of dissolution structures and relative depths


of penetration obtained when injecting the same volume of fluid
at different Damkhler numbers.

Fig. 10 Effect of number of wormholes on skin evolution for


constant injection rate.

6
5

(5) Uniform dissolution


(4) Ramified wormholes
(3) Dominant wormholes
(2) Conical wormholes
(1) Face Dissolution

1,000

Skin [dimensionless]

Wormhole
[#/ft2]]
Wormhole density [#/ft

10,000

100
(5)
(4)

10

4
3
2
HCl

1
HFc
0

HAc

-1
Emulsified HCl

(3)

-2

(2)
1 (1)
0

-3
2

10

12

Depth of penetration [in]

Fig. 9 Dependence of wormhole density on depth of


penetration for a variety of dissolution structures.

10

Volume injected [bbl]

Fig. 11 Comparison of skin evolution for a variety of


stimulation fluids at 200F.

C.N. FREDD

Normalized optimum injection rate


[dimensionless]

60
50
Typical fluid loss

40
30
Low fluid loss

20
10

8
HCl
HFc

HAc

Emulsified HCl

0
50

0
Depth of penetration

HCl

HFc

HAc

Emulsified HCl

150

250

350

450

Temperature [F]

Fig. 12 Effect of fluid loss on optimum injection strategies for


effective wormhole formation with HCl at 200F.

Normalized optimum injection rate


[dimensionless]

SPE 59537

Normalized optimum injection rate


[dimensionless]

14

Changing fluid
type at constant
injection rate

0
Depth of penetration
Fig. 13 Optimum injection strategies for various acid systems
at 200F. (The shaded box represents typical injection rates used
in matrix acidizing treatments.)

Fig. 14 Effect of temperature on the optimum injection rate


required to achieve wormhole penetration of 6 in. with various
acid systems. (The shaded box represents typical injection rates
used in matrix acidizing treatments.)

Anda mungkin juga menyukai