Sunteți pe pagina 1din 10

SPE 54719

A Fundamentally New Model of Acid Wormholing in Carbonates


Rick Gdanski, SPE, Halliburton Energy Services, Inc.
Copyright 1999, Society of Petroleum Engineers, Inc.
This paper was prepared for presentation at the 1999 European Formation Damage Confer-
ence held in The Hague, The Netherlands, May 31-June 1.
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. Permission to copy is restricted to an abstract of not more than 300
words. Illustrations may not be copied. The abstract should contain conspicuous acknowledg-
ment of where and by whom the paper is presented. Write Librarian, SPE, P.O. Box 833836,
Richardson, TX 75083-3836, U.S.A, fax 01-972-952-9435.
References at the end of the paper.
Abstract
A new theoretical model has been developed to describe the
chemical reactions of acid in porous carbonate media. For the
past 20 years, the three major unanswered questions regarding
wormholing when acid is pumped into carbonate formations
have been the following: (1) How many dominant wormholes are
created? (2) What is the spatial distribution of those dominant
wormholes along the wellbore? (3) What is the leakoff profile
from the dominant wormholes under radial flow conditions?
This paper presents a model that proposes answers to these three
basic questions. Once these questions are answered, the reaction
of acid in the matrix and the interaction of wormhole develop-
ment become straightforward. The new model requires a major
paradigm shift in the understanding of matrix carbonate acidizing
and the variables that control wormhole growth. Parametric
studies that were conducted with the new theory are presented. It
was found that wormhole length is predominantly controlled by
matrix porosity and volume of acid pumpednot by reactivity.
It was found that wormhole diameter is predominantly con-
trolled by reactivity and contact time. The new theory confirms
classically held guidelines for matrix acidizing of carbonates,
and gives insight on how to improve matrix carbonate acidizing
treatments. The new model also accommodates the effects of
permeability anisotropy caused by natural fracturing or layering
effects.
Introduction
In 1979, SPE published Monograph Volume 6 of the Henry L.
Doherty Series Acidizing Fundamentals which was coauthored
by Bert Williams, John Gidley, and Robert Schechter.
1
The
monograph is an excellent reference about matrix acidizing
sandstones, fracture acidizing carbonates, and matrix acidizing
carbonates. Subsequent improvements to fracture acidizing
theory include balancing mass transport and surface reaction for
calculation of acid spending,
2,3
incorporating heat of reaction,
4
measuring and using energy of activations for reactivity profiles
as a function of temperature,
5
improving fracture geometry
calculations for fracture acidizing with multiple stages,
6
deter-
mining reactivity for slow-reacting gelled acids,
7
and introduc-
ing an improved method for using laboratory-measured conduc-
tivity data in a fracture acidizing simulator.
8
Matrix acidizing of carbonates is also covered in the acidizing
monograph. A method is given for calculating the spending of
acid down a dominant wormhole in either turbulent or laminar
flow. Hand calculations of acid spending lengths can be per-
formed either with or without fluid leakoff. Unfortunately, the
following three fundamental questions that have prevented use
of the published concepts have remained unanswered for the
past 20 years:
1. How many dominant wormholes are generated?
2. What is the spatial distribution of these dominant worm-
holes?
3. What is the leakoff profile from the dominant wormholes?
A new theory presented in this paper proposes to answer
these three fundamental questions.
Current State-Of-Affairs
At five research centers throughout the world, significant stud-
ies of acid wormholing in carbonates have been conducted. The
two most prolific places have been the University of Michigan
in Ann Arbor under the direction of Scott Fogler and the
University of Texas in Austin under the direction of Dan Hill.
The other locations are the Mining University Leoben with
Thomas Frick under the direction of Michael Economides, the
Institute Franais du Ptrole with Brigitte Bazin, and the
Halliburton Energy Services, Inc. (HES) European Research
Center (ERC) with Martin Buijse.
Scott Fogler and graduate students at the University of
Michigan have conducted many experiments and have visual-
ized the wormhole patterns with metal castings
9
and later with
2 SPE 54719 A FUNDAMENTALLY NEW MODEL OF ACID WORMHOLING IN CARBONATES
neutron radiography.
10
They conducted linear-flow experiments
at various flow rates with carbonates of different reactivities,
and modeled the general pattern of the wormholes with a
network model.
11
Unfortunately, such a computer model has
great difficulty in scaling to the real world. Nonetheless, they
demonstrated that the wormholing patterns were a function of
the Damkhler number, which is essentially the ratio of reactiv-
ity to mass transport. They recently introduced a generalized
Damkhler number and an optimum Damkhler number for
most efficient generation of wormholes in carbonates.
12
The
concept is apparently useful for both slowly reacting organic
acids as well as HCl on either limestones or dolomites. Unfor-
tunately, the concept has been developed through the use of
linear-flow experiments, and it is uncertain how it will scale to
radial flow.
Dan Hill and his graduate students at the University of
Texas conducted a large number of linear-flow experiments and
have attempted to define an optimum flow rate for acidizing.
13 ,14
This approach is very similar to that taken at the University of
Michigan. Recently, the University of Texas group has intro-
duced some new mathematics and charts for calculating the
optimum injection rate for initiating efficient wormholing at the
wellbore. However, their predictions for the optimum injection
rate for acidizing carbonate reservoirs still seem unrealistically
low, and the method is cumbersome.
Thomas Frick, at the Mining University Leoben under the
direction of Michael Economides, published work a few years
ago in which fractal mathematics was used to describe the
branching nature of the wormhole patterns observed in radial
acidizing experiments of a quarried carbonate.
15 ,16
Unfortu-
nately, such calculations were exceedingly time-intensive and
required weeks of CPU time on a powerful workstation. The
results of these calculations were still not general enough to
allow use in other carbonate reservoirs of differing lithology and
reactivity. Later, Behdokht Mostofizadeh, while at the Mining
University Leoben under the direction of Michael Economides,
published an approach to predict the optimum injection rate
from radial acidizing experiments.
17
Brigitte Bazin at Institute Franais du Ptrole has investi-
gated the implications of linear wormholing on the leakoff
properties of acid during fracture acidizing treatments.
18 ,19
While the work is interesting, it is not clear that there is a ready
application in either fracture acidizing or matrix acidizing of
carbonates.
Martin Buijse at the HES ERC has been performing both
linear and radial flow experiments in an attempt to understand
the difference of wormholing behavior in linear flow vs. radial
flow. In addition, he recently published an approach for calcu-
lating acid penetration distances down a wormhole under lami-
nar flow conditions.
20
Interestingly, the recent work indicates
that the classic pore volumes of acid to breakthrough in linear-
flow experiments may be an artifact of the core diameter.
Therefore, the apparent artifact suggests that the linear experi-
ments may not be capable of contributing much more to the
understanding of wormhole growth in radial flow.
Grard Daccord with Dowell-Schlumberger in Saint Etienne,
France has published several papers on the wormhole patterns
formed by the dissolution of plaster with water.
21 ,22 ,23
His
important work, though seemingly unrelated to acid wormholing
in carbonates, provided fundamental knowledge for the devel-
opment of the new theory presented in this report.
In the nearly 20 years since the SPE acidizing monograph
was published, a great deal of work has been performed, but the
three fundamental questions have remained unanswered.
The Breakthrough
In the mid-1980s, the author attended a meeting of the Indus-
trial Affiliates Program at the University of Michigan under the
direction of Scott Fogler. Results of carbonate wormholing
studies were presented, and pictures of some radial flow disso-
lution experiments were displayed. The experiments by Daccord
were actually dissolution patterns created by water injection into
plaster casts. The pictures displayed an amazingly high degree
of symmetry that was clearly evident in the wormholing pat-
terns. Daccord published a picture (Fig. 1, Page 9) in SPEPE in
1989.
23
It was only a side view, and the full symmetry was not
as obvious. Scott Fogler also had copies of top views that were
not published.
The high degree of symmetry in the radial flow dissolution
experiments was profoundly significant. Symmetry occurs in
highly coupled systems, and its presence can dramatically affect
the ease of understanding and modeling of a particular phenom-
enon. For example, organic and inorganic chemists use symme-
try to predict certain reactivity characteristics. The existence of
symmetry in electron wave functions of certain reactive mol-
ecules allows one to predict whether certain types of reactions
will proceed. When such symmetry is present, it is not necessary
to spend the months of CPU time that might be required to
calculate the reaction potential. Rather, one can look at the
structure drawn on paper and say this will proceed or this will
not proceed. Likewise, the existence of symmetry and/or the
breaking of symmetry in the study of high-energy nuclear
physics phenomena can have an enormous impact on the theo-
ries that physicists may accept or dismiss. In fact, in grand
unified theories on the nature of the universe, the breaking of
symmetry gives us gravity vs. electromagnetic force vs. strong
nuclear force, etc. Similarly (but certainly much less signifi-
cantly), the symmetry in the dominant wormholing patterns for
radial flow dissolution allows for significant insight and simpli-
fication of the phenomenon.
The radial dissolution experiments of Daccord consisted of
water being flowed from a wellbore installed in a hardened
plaster cylinder. The observed symmetry existed both radially
from the center and linearly along the length of the wellbore.
Dominant wormholes occurred as sets of either five or six in
a single horizontal plane. The wormholes in sets of six had a
SPE 54719 3 R. GDANSKI
radial separation of 60 from each other. Such a distribution
essentially formed a series of equilateral triangles. Thus, the
distance between the tips of the separate wormholes in each set
was the same as the length of the wormholes. These horizontal
sets occurred in a periodic fashion along the vertical length of
the wellbore but had different radial penetrations. A number of
sets of wormholes approximately 1 in. in radial length were
separated from each other by approximately 1 in. Inspection
showed that every other set had ceased to grow, and that the
remainder continued to grow until their length was equal to the
separation distance between them (approximately 2 in.). Again,
about half of those sets stopped growing while the remaining
sets continued to grow until one wormhole broke through the
edge of the plaster cylinder.
The presence of the symmetry indicates that all the domi-
nant wormholes must know about each otherthe
wormholing process is a highly coupled phenomenon, and
dominant wormholes do not form randomly from each other in
radial flow. The mechanism for coupling is through the pressure
field induced by matrix flow of fluid through the porous media.
This includes interferences between developing dominant worm-
holes through the matrix flow of both displaced pore fluids and
invading spent fluids. Minimizing these interferences requires
that the dominant wormholes stay as far apart as possible. As a
consequence, the development of multiple wormholes in radial
flow must be symmetrical both radially and vertically. In addi-
tion, the wormholes must also observe the influences of Darcys
law for matrix radial flow. Most investigators assume that the
direction of fluid flow through the matrix is governed by the
developing wormhole pattern. As a result, they focus on the
physics of the wormhole growth and ignore the matrix itself.
The breakthrough in thinking is that the developing wormhole
pattern is governed by fluid flow through the matrix. Once
this concept is accepted, the diameter and length of the develop-
ing dominant wormholes are readily calculated from acid reac-
tivity and spending procedures currently used in advanced
fracture-acidizing theory.
The New Theory
The new theory introduced in this paper is that wormholing of
acid in carbonates is highly symmetrical when it is performed on
the scale of field acid treatments. In addition, the secondary
shape of this symmetry is governed by the native formation
permeability. This symmetry was experimentally observable
with radial flow of water through plaster casts because of the low
solubility of plaster in water and the fairly homogeneous nature
of hardened plaster, thereby minimizing end effects. In the
few radial carbonate flow tests conducted at the ERC, the high
solubility of carbonate in HCl, the generation of a CO
2
gas
phase, and the fairly nonhomogeneous nature of carbonate cores
prevented the high degree of symmetry from forming before
wormhole breakthrough occurred. Therefore, end effects were
a dominant part of the wormholing pattern. However, a close
inspection of casts from these radial HCl flow experiments
indicated that the symmetrical acidizing was initially trying to
form.
24
Such symmetry was never, and could never, be observed
in linear cores because of the fundamental difference between
linear and radial flow.
The assumptions for the new theory follow:
1. The spatial distribution of the dominant wormhole pattern
is governed by the native permeability of the carbonate
formation.
2. The fluid velocity through the wormhole pattern converges
at the tip to the fluid velocity calculated by uniform matrix
radial flow as if no wormhole pattern existed. (This condi-
tion is essentially the leading edge boundary condition, and
it is a fundamental assumption for calculating the fluid
leakoff profile from the developing dominant wormholes.)
3. Dominant wormholes form as sets of six in an x-y plane
perpendicular to the wellbore and independent of perfora-
tion density.
4. The radial separation of the wormholes in each set is 60.
5. Two opposing dominant wormholes in each set align them-
selves with the direction of highest permeability, which is
assumed to be in the x direction.
6. The tip length undergoing leakoff for each dominant worm-
hole is the same as one-half the length of the 60 leading
edge arc assigned to that wormhole.
7. The remaining wormhole length (closest to the wellbore) is
simply a conduit for transporting acid from the wellbore; it
does not undergo leakoff but does undergo reaction and
growth by dissolution.
8. The secondary shape of the dominant wormhole pattern in
each set is governed by the permeability contrast between
the x and y directions.
9. The frequency of the sets of wormholes along the z direc-
tion parallel to the wellbore is governed by the permeability
contrast between the x and z directions.
10. The separation of growing wormhole sets cannot be closer
than the dominant wormhole length when the x and z
permeabilities are the same. (This is essentially how one
determines when to reduce the number of developing worm-
hole sets.)
These assumptions are rather technical but can help provide
the answers to the three unanswered questions. The assumptions
also allow a straightforward programming of the model. In
addition, the reaction of the acid down the wormholes is calcu-
lated according to the balanced acid reactivity/mass transport
approach of Roberts and Guin.
2
The mass transport coefficients
for turbulent and transitional flow regimes in tubes were ob-
tained from subroutines in an HES fracture acidizing program.
These coefficients had been obtained from literature sources
cited by Roberts and Guin. The fracture acidizing program uses
mass transport coefficients for tubes and converts them for use
4 SPE 54719 A FUNDAMENTALLY NEW MODEL OF ACID WORMHOLING IN CARBONATES
in parallel plates (fractures). The mass transport for laminar flow
through reactive tubes was obtained from Levich.
25
General Algorithm
A computer model was written based on the assumptions listed
above. The first step was to determine the initial number of
wormhole sets. For a homogeneous formation, the number of
sets was chosen to be a value of (2
i
1), where i is an integer, that
would give an even distribution of sets along the zone height
with a separation of 18 in. or less. Each wormhole set consisted
of six dominant wormholes at a radial separation of 60 from
each other. The wormholes were initially defined as being 6 in.
long with a radius at the wellbore of 0.03 in. and a radius at the
tip of 0.01 in. The matrix pore throats ahead of the growing
dominant wormholes were also assumed to have an initial radius
of 0.01 in.
An acid volume was then injected into the formation. The
acid volume was calculated to be the incremental volume
required to fill the matrix porosity for the next radial length step
ahead of the growing wormhole sets. The radial length step was
set at 0.20 in. Fluid distribution per wormhole was calculated to
be the volume step divided by the total number of dominant
wormholes.
Fluid leakoff was calculated as the injected fluid was moved
down the length of the dominant wormholes. In a homogeneous
formation, the length of each dominant wormhole experiencing
leakoff was calculated as one-half the arc length associated with
each wormhole. For example, a wormhole set having dominant
wormholes extending 5 ft from a 6-in. diameter wellbore would
have a circumference of 33 ft. Each wormhole has an associated
arc length of 5.5 ft. Leakoff from each dominant wormhole
would be calculated along the leading length of 2.75 ft, or 55%
of the leading length of each dominant wormhole. No leakoff
would be allowed for the first 45% of the wormhole length.
Enough acid leakoff would be calculated to leave just enough
fluid at the tip to fill the associated radial increment ahead of the
wormhole tip. This leakoff procedure ensured compliance with
Assumption 2 and is fundamental to the new theory. It is the
method by which a homogeneous pressure field is established
and through which coupling of the wormhole network be-
comes symmetrical.
Acid spending down the length of the wormholes was then
calculated. The procedure was a finite element spending routine
that used the balanced surface reactivity/mass transport ap-
proach of Roberts and Guin.
2,3
New wormhole dimensions were
calculated, and the process was repeated.
The number of sets of wormholes was reduced by about
one-half when the length of the dominant wormholes became
the same as the vertical separation of each wormhole set.
However, the number of wormhole sets was always kept at some
multiple of (2
i
1).
Results and Discussion
General Isotropic Example. Visualization of the new theory is
critical for understanding its simplicity as well as its impact and
utility. An example will be given where the dominant wormhole
pattern has been calculated and visualized. The example is for a
10-ft zone acidized with 15% HCl at a surface pumping rate of
0.1 bbl/min/ft of zone or 1 bbl/min. The treatment volume was
also at the conventional value of 100 gal/ft of zone or 1,000 gal.
The reactivity was assumed to be high at 150F in limestone with
10% porosity and uniform permeability. Table 1 (Page 8) lists
the specific details.
The results of the wormhole calculations are shown in Figs.
2 through 7 (Pages 9 and 10). Figs. 2 and 3 show the wormholing
pattern from both a side view and a top view at the end of the
treatment. The vertical and radial symmetry is very similar to the
symmetry that was observed by the author several years ago at
the University of Michigan. This symmetry has now been
incorporated into the modeling of the new wormholing theory.
Fig. 2 shows that at the end of the treatment there was only one
set of growing wormholes. Initially, there were seven sets, but
four of them died out within 1 minute and two more died out
within 4 minutes of starting the treatment. This left one domi-
nant wormhole set for a total of six wormholes taking the
remainder of the acid for the final 20 minutes of pumping.
It is also appropriate to use Fig. 3 (Page 9) to define primary
and secondary dominant wormholes. Figs. 4 through 7 (Page 10)
show various calculated values for the dominant wormholes. In
Fig. 3, the primary dominant wormholes are the two wormholes
aligned with the x-axis. The secondary dominant wormholes are
the four wormholes aligned at a 60 angle from the x-axis. When
the permeability in the x and y directions are the same, the
calculated values for the primary and secondary dominant
wormholes are also the same. However, if a permeability con-
trast exists between the x and y directions, the definition of
primary and secondary wormholes becomes important, since
the calculated values will be different.
Fig. 4 (Page 10) shows the radius of the wormholes as a
function of length in the final growing wormhole set. Fig. 5
(Page 10) shows the acid spending profile along the length of the
dominant wormholes. The diameter of the wormholes was at
least 0.25 in. (>0.125-in. radius) for at least 60 in. (5 ft) from the
wellbore and was over 1 in. in diameter for almost 4 ft from the
wellbore. In addition, the bulk of the acid spending was in the
leading 20% of the wormhole tip because of the longer residence
time in the developing wormholes as fluid leakoff occurred. The
smaller wormhole diameters near the tip also allowed more
efficient acid spending since they had a higher surface-area-to-
acid-volume ratio. Even though there was only one dominant
wormhole set at the end of the treatment, their large size would
still have allowed very effective production improvement.
Linear core testing has established that an optimum
Damkhler number exists for creating wormholes most effi-
ciently. This number is normally used in graphical form as its
inverse, 1/Da. The optimum 1/Da for high-reactivity limestone
SPE 54719 5 R. GDANSKI
is thought to range from approximately 1 to 10. In Fig. 6 (Page
10), the value of 1/Da fell below 1 after about 5 minutes of
pumping. Keeping that value mid-range between 1 and 10
would have required increasing the pumping rate from about 1
bbl/min at the start of the treatment to about 3 to 5 bbl/min at the
end of the treatment, which is, of course, a common practice for
matrix acidizing limestones. Thus, the common practice of
pumping at 0.1 bbl/min/ft and increasing the pumping rate as
surface treating pressure allows has kept our industry in a very
reasonable range for effective matrix acidizing of carbonates.
While this is easy to accomplish in small zones, it is much more
difficult to accomplish in long horizontal wells.
In the linear-flow experiments, failure to stay at the opti-
mum 1/Da results in multiple pore volumes of acid being
required to effectively propagate wormholes to a length of 6 in.
In the new radial model, radial leakoff along the length of the
wormholes resulted in continually keeping fresh acid near the
tip of the wormhole. The result is much more efficient acidizing
in radial flow than in linear flow.
Fig. 7 (Page 10) shows the Reynolds Number calculated at
the entrance of the growing wormholes as a function of time.
This calculation has significant error in it during the first few
minutes of the treatment. The new model assumes only a certain
number of holes will take fluid at the start of the treatment, while
in reality, there could be thousands more taking fluid. However,
within 1 minute of the start of acid injection, the number of holes
taking fluid had dramatically reduced to just four sets of six, or
24 dominant wormholes. From then on, the calculation of the
Reynolds number is reasonable. Therefore, the acid entered the
growing wormholes in deep turbulent flow. At the end of the
treatment the acid was still entering each wormhole at a Reynolds
number of over 20,000, which was not even close to laminar
flow, in which the Reynolds number is less than 1,800.
Parametric Study. The properties of the new model were
studied by examination of a few typical parameters, including
formation porosity, carbonate reactivity, treatment pumping
rate, and acid volume (Table 2, Page 8). The conditions were
similar to the previous example, except the thickness was
increased to 50 ft to increase the number of wormhole sets
during most of the treatment designs. However, the normalized
fluid volume and pump rate remained the same at 100 gal/ft and
0.1 bbl/min/ft of zone.
Porosity Effects. A series of calculations was performed to
demonstrate the effect of formation porosity on wormhole
penetration. For example, chalk formations tend to have porosi-
ties in the range of 40 to 60%, while dense limestones tend to
have porosities less than 10%. It was assumed that all of the
matrix porosity was accessible to acid flow. The reference data
for the calculations are listed in Table 2 (Page 8). The porosity
was varied from 2 to 50%.
The results of the wormhole calculations are shown in Fig.
8 (Page 10). The impact of formation porosity is a direct
consequence of the requirements of fluid invasion. With a
porosity of 50%, the wormhole penetration was less than 3 ft,
even with 100 gal/ft of acid because the invading acid con-
formed to Darcys law for matrix radial flow. Therefore, the
invading acid must fill all the porosity in its path and displace the
resident fluids. The net effect is a large number of short worm-
holesthere were 15 sets of dominant wormholes across the 50-
ft interval for a total of 90 dominant wormholes. By contrast, the
2% porosity calculation could represent a dense limestone with
a high degree of uniform natural fracturing. The effective
secondary porosity of 2% might be all that is readily accessible.
In such a case, there is little place for the invading acid to go, so
it must invade deep into the formation. At the end of the
treatment, there were only three sets of dominant wormholes
across the 50-ft interval for a total of 18 dominant wormholes.
It is clear that formation porosity is an important factor for
treatment design.
Reaction Rate Effects. A series of calculations was performed
to demonstrate the effect of formation reactivity on wormhole
penetration. For example, dolomite formations at cool tempera-
tures tend to have lower reactivities than limestones. The linear-
flow experiments at the universities have shown that slow-
reactivity formations tend not to generate dominant wormholes
easily. The reference data for the calculations are listed in Table
2 (Page 8). The reaction rate constant was varied from 1.0E-2 to
3.0E-5.
The results of the wormhole calculations are shown in Fig.
9 (Page 10). As the reaction rate dropped below 3.0E-4, the
wormhole diameter was much smaller even though the worm-
hole length remained unchanged. The nominal 1/Da (inverse of
the Damkhler number) rose above 50 as the reaction rate
dropped below 3.0E-4, and it easily exceeded 500 as the reaction
rate dropped to 3.0E-5. The new wormholing theory is therefore
in qualitative agreement with the predictions for dominant
wormhole diameters from the Damkhler number approach. In
addition, this example shows that reactivity plays an important
role on wormhole diameter but not on penetration distance.
Pumping Rate Effects. A series of calculations was performed
to demonstrate the effect of pumping rate on wormhole penetra-
tion in high-reactivity carbonates. The linear-flow experiments
at the universities have shown that slow injection rates tend
toward face dissolution while high injection rates tend to make
thin, dominant wormholes. The reference data for the calcula-
tions are listed in Table 2 (Page 8). The surface pumping rate was
varied from 0.05 to 500 bbl/min. The corresponding normalized
pumping rates were 0.001 to 10 bbl/min/ft of zone. The corre-
sponding treatment times were 40 hours to 14 seconds.
The results of the wormhole calculations are shown in Fig.
10 (Page 10). As the pumping rate dropped to 0.05 bbl/min
6 SPE 54719 A FUNDAMENTALLY NEW MODEL OF ACID WORMHOLING IN CARBONATES
(0.001 bbl/min/ft of zone), the wormhole radius at the wellbore
became quite large. The nominal 1/Da during this treatment was
in the range of 0.007 to 0.07 and is clearly in the range of face
dissolution as predicted from the Damkhler number approach.
At a pumping rate of 500 bbl/min (10 bbl/min/ft), the wormhole
diameter was much smaller, as would be expected. Therefore,
the new wormholing theory is still in qualitative agreement with
the Damkhler number approach. However, it is clear that there
is a great deal of latitude for an acceptable injection rate for
acidizing.
Acid Volume Effects. A series of calculations was performed
to demonstrate the effect of acid volume on wormhole penetra-
tion in high-reactivity carbonates. Common questions when
treating carbonates include How much acid should be pumped?
and Does doubling the acid volume get twice the wormhole
length? The reference data for the calculations are listed in
Table 2 (Page 8). The acid volume was varied from 500 to
50,000 gal. The corresponding normalized volumes were 10 to
1,000 gal/ft. The corresponding treatment times were 2 minutes
to 4 hours.
The results of the wormhole calculations are shown in Fig.
11 (Page 8). At 10 gal/ft of zone (500 gal), the acid easily
penetrates about a foot from the wellbore in a 10%-porosity
carbonate. Pumping the conventional volume of 100 gal/ft
(5,000 gal) gives a wormhole penetration of about 5 ft. Dou-
bling this value to 200 gal/ft increases the wormhole penetration
by only 50% to about 8 ft. Creating a wormhole pattern to a
distance of 20 ft would require over 1,000 gal/ft or 50,000 gal
of acid in a 50-ft zone. This observation suggests that designs for
improving wormhole penetration might require some kind of
fluid-loss control measure to prevent the acid from invading the
entire porosity.
Implications
The preceding examples show the potential utility of the new
wormholing theory as incorporated in the computer model. The
model was written with the expectation that a spent acid front
would precede the wormhole development front as suggested
by the multiple pore volumes of acid required in laboratory
linear-flow experiments. In addition, the spending routine was
originally written assuming laminar flow in the wormholes.
However, as the model was studied for its behavior, it was found
that the flow in the wormholes was usually turbulent and live
acid stayed very close to the fluid invasion front. This behavior
required incorporation of turbulent flow spending and a closer
examination of fluid leakoff. In addition, it was initially puz-
zling that live acid could penetrate well over 30 ft into a high-
reactivity matrix. This observation put the spending routine in
question. Therefore, hand calculations performed by the author
several years ago using the acidizing monograph mathematics
and charts were reviewed. Indeed, HCl is quite capable of
traveling down a wormhole for over 100 ft in a high-reactivity
carbonate, depending on conditions such as wormhole diameter,
fluid velocity, and fluid leakoff. In that regard, the acidizing
monograph and the new wormholing theory are in agreement.
With the original expectation that acid spending would be
the main influence on wormhole penetration distance, the com-
puter model was initially written to maximize accuracy of acid
spending. Unfortunately, with low-reactivity carbonates, the
CPU requirements on a 486 class PC easily exceed 12 hours.
Once it became clear that the main influence on wormhole
penetration distance was simply fluid invasion, the computer
model was modified to make fluid invasion and leakoff more
accurate and let acid spending down the wormholes become
secondary. The result was that most calculations took about 2 to
5 minutes of CPU time with no significant loss of accuracy in
acid spending or wormhole growth.
The findings from the model indicate several things. First,
reactivity in most practical matrix treatments generates worm-
hole diameter. Specifically, live acid penetration to 20 ft from
the wellbore is governed by pumping enough acid to fill the
accessible matrix porosity to a radial distance of 20 ft, regardless
of reactivity. If the carbonate is of low reactivity, then the
wormholes will be exceedingly small in diameter, and produc-
tion increase will be disappointing. If the carbonate is of moder-
ate to high reactivity, then the wormholes will be of moderate to
large diameter, and production increases will be as expected. In
both cases, the live acid will have penetrated the 20 ft, even
though the same rock-dissolving power is used. The difference
is that low- reactivity carbonates are dissolved uniformly through-
out the matrix while high-reactivity carbonates are dissolved
more in the dominant wormholes than in the matrix.
Second, since wormhole length is governed simply by fluid
invasion distance, higher acid concentrations only affect worm-
hole diameters and not wormhole length. This consideration
might be important for a low-reactivity carbonate where worm-
hole diameter development is slow. However, for a high-reactiv-
ity carbonate there would be no advantage to using 28% HCl
over 15% HCl because the wormhole diameter would already be
large enough with 15% HCl.
Third, the most important factor in determining acid pen-
etration is the matrix invasion distance of the acid. The forma-
tion porosity and the volume of acid pumped into the formation
control the invasion distance, so it becomes important to know
both the formation porosity and the accessible porosity to flow.
The model can be calibrated by adjusting the accessible porosity
to match poststimulation skins. The impact of porosity can
finally help explain why high-porosity chalks can be so difficult
to stimulate to the same level as classical carbonates with
porosities of 10 to 15%.
Finally, improved wormhole penetration distance will likely
depend on methods that prevent acid from filling the entire
matrix porosity, thereby reducing the effective or accessible
porosity for the treatment. This could be accomplished with
fluid-loss control measures such as oil-soluble resin, gel diverters,
or foam diversion.
SPE 54719 7 R. GDANSKI
Model Calibration
As a test of the new theory, test results were matched with matrix
acidizing treatments. The literature reports treating and forma-
tion conditions for a highly diverted acid treatment in a carbon-
ate formation offshore from Qatar.
26
Table 3 (Page 8) reports
the formation and treatment data for two of the wells. The
postacid treatment skins and model calibration information are
included as well.
The model was calibrated based on the assumption that the
permeability improvement from acid was at least 50 times
greater than native permeability to a radial distance equal to the
acid penetration distance. A skin was then calculated and com-
pared to the field treatment. The average acid penetration
distance for each field treatment was determined by application
of Hawkins equation. The porosity in the model was then
adjusted (reduced) until the calculated negative skin matched
the pressure buildup skin determined after the acid treatment.
Table 3 (Page 8) shows that the accessible porosity of the
carbonates was significantly lower than the matrix porosity. In
other words, the formation required significantly less than 1
pore volume (PV) of acid to stimulate to the necessary radial
distance. The match showed that about 0.4 PV of acid was
required in the 100-md formation and about 0.3 PV was required
in the 5-md formation. This observation contrasts with classical
views of high-reactivity carbonates requiring multiple pore
volumes of HCl to achieve a given wormhole penetration
distance.
Conclusions
1. A new carbonate matrix acidizing wormholing model has
been proposed. This model takes advantage of the experi-
mentally observed symmetry of wormholing under radial
flow conditions.
2. Wormhole penetration distances are predominantly con-
trolled by formation porosity and acid volume.
3. Formation reactivity and contact time with the acid pre-
dominantly control wormhole diameters and acidized per-
meability.
4. The new theory confirms classically held guidelines for
matrix acidizing of carbonates.
5. Calibration of the model with field treatments indicates
only fractional pore volumes of acid are required to achieve
a given stimulation distancenot multiple pore volumes.
References
1. Williams, B.B., Gidley, J.L., and Schechter, R.S.: Acidizing
Fundamentals, Monograph Volume 6, Henry L. Doherty Series,
SPE-AIME, Dallas (1979).
2. Roberts, L.D. and Guin, J.A.: A New Method for Predicting Acid
Penetration Distance, SPEJ (August 1975) 277-286.
3. Gdanski, R.D. and van Domelen, M.S.: Slaying the Myth of
Infinite Reactivity of Carbonates, paper SPE 50730 presented at
the 1999 International Symposium on Oilfield Chemistry, Hous-
ton, TX Feb. 16-19.
4. Lee, M.H. and Roberts, L.D.: Effect of Heat of Reaction on
Temperature Distribution and Acid Penetration in a Fracture,
SPEJ (December 1980) 501-507.
5. Anderson, M.S.: Reactivity of San Andres Dolomite, SPEPE
(May 1991) 227-232.
6. Lee, W.S.: Geometry Determination for Multi-Stage Acidizing
Treatments With or Without Viscous Preflush, paper SPE 14515
presented at the 1985 Eastern Regional Meeting, Morgantown,
WV, Nov. 6-8.
7. Gdanski, R.D. and Norman, L.R.: The Effect of Filterable Solids
on Acid Reaction Rates, SPEPE (March 1986) 111-116.
8. Gdanski, R.D. and Lee, W.S.: On the Design of Fracture Acidizing
Treatments, paper SPE 18885 presented at the 1989 Production
Operations Symposium, Oklahoma City, OK, Mar. 13-14.
9. Hoefner, M.L. and Fogler, H.S.: Pore Evolution and Channel
Formation During Flow and Reaction in Porous Media, AIChEJ
34 1 (1988) 45-54.
10. Jasti, J.K. and Fogler, H.S.: Application of Neutron Radiography
to Image Flow Phenomena in Porous Media,: AIChEJ 38 4 (1991)
19-26.
11. Hoefner, M.L. and Fogler, H.S.: Fluid-Velocity and Reaction-
Rate Effects During Carbonate Acidizing: Application of Net-
work Model, SPEPE (February 1989) 56-62.
12. Fredd, C.M., Tjia, R., and Fogler, H.S.: The Existence of an
Optimum Damkhler Number for Matrix Stimulation of Carbon-
ate Formations, paper SPE 38167 presented at the 1997 European
Formation Damage Conference, The Hague, The Netherlands,
Jun. 2-3.
13. Wang, Y., Hill, A.D., and Schechter, R.S.: The Optimum Injec-
tion Rate for Matrix Acidizing of Carbonate Formations, paper
SPE 26578 presented at the 1993 Annual Technical Conference
and Exhibition, Houston, Oct. 3-6.
14. Huang, T., Hill, A.D., and Schechter, R.S.: Reaction Rate and
Fluid Loss: The Keys to Wormhole Initiation and Propagation in
Carbonate Acidizing, paper SPE 37312 presented at the 1997
International Symposium on Oilfield Chemistry, Houston, Feb.
18-21.
15. Frick, T.P., Krmayr, M., and Economides, M.J.: Modeling of
Fractal Patterns in Matrix Acidizing and their Impact on Well
Performance, SPEPF (February 1994) 61-68.
16. Frick, T.P., Mostofizadeh, B., and Economides, M.J.: Analysis
of Radial Core Experiments for Hydrochloric Acid Interaction
With Limestones, paper SPE 27402 presented at the 1994 Inter-
national Symposium on Formation Damage Control, Lafayette,
Feb. 7-10.
17. Mostofizadeh, B. and Economides, M.J.: Optimum Injection
Rate From Radial Acidizing Experiments, paper SPE 28547
presented at the 1994 Annual Technical Conference and Exhibi-
tion, New Orleans, Sept. 25-28.
18. Bazin, B., Roque, C., and Boutca, M.: A Laboratory Evaluation
of Acid Propagation in Relation to Acid Fracturing: Results and
Interpretation, paper SPE 30085 presented at the 1995 European
Formation Damage Conference, The Hague, The Netherlands,
May 15-16.
19. Bazin, B., Roque, C., Chauveteau, G., and Boutca, M.: Acid
Filtration in Dynamic Conditions to Mimic Fluid Loss in Acid
Fracturing, paper SPE 38168 presented at the 1997 European
Formation Damage Conference, The Hague, The Netherlands,
Jun. 2-3, 1997.
8 SPE 54719 A FUNDAMENTALLY NEW MODEL OF ACID WORMHOLING IN CARBONATES
20. Buijse, M.A.: Understanding Wormholing Mechanisms Can
Improve Acid Treatments in Carbonate Formations, paper SPE
38166 presented at the 1997 European Formation Damage Con-
ference, The Hague, The Netherlands, Jun. 2-3.
21. Daccord, G.: Chemical Dissolution of a Porous Medium by a
Reactive Fluid, Phys. Rev. Lett. (1987) 58, 479-482.
22 . Daccord, G. and Lenormand, R.: Fractal Patterns from Chemical
Dissolution, Nature (1987) 325, 41-43.
23. Daccord, G., Touboul, E., and Lenormand, R.: Carbonate
Acidizing: Toward a Quantitative Model of the Wormholing
Phenomenon, SPEPE (February 1989) 63-68.
24 . MaGee, J., Buijse, M.A., and Pongratz, R.: Method for Effective
Fluid Diversion When Performing a Matrix Acid Stimulation in
Carbonate Formations, paper SPE 37736 presented at the 1997
Middle East Oil Show, Bahrain, Mar. 15-18.
25. Levich, V.G.: Physicochemical Hydrodynamics, Prentice-Hall
Inc., Englewood Cliffs, NJ (1962).
26. MaGee, J., Buijse, M.A. ,and Pongratz, R.: Method for Effective
Fluid Diversion when Performing a Matrix Acid Stimulation in
Carbonate Formations, paper SPE 37736 presented at the 1997
Middle East Oil Show, Bahrain, Mar. 15-18.
Temperature 150F
Zone Thickness 10 ft
x Permeability 1 md
y Permeability 1 md
z Permeability 1 md
Porosity 10%
Reaction-Rate Constant 1.0E-2
Reaction Order 0.5
Surface Pump Rate 1 bbl/min
HCl Concentration 15%
Volume of HCl 1,000 gal
Total Treatment Time 23.8 min
Normalized Pump Rate 0.1 bbl/min/ft
Normalized Volume 100 gal/ft
Treatment Design
Table 1General Isotropic Example
Formation Characteristics
Temperature 150F
Zone Thickness 50 ft
x Permeability 1 md
y Permeability 1 md
z Permeability 1 md
Porosity 10%
Reaction Rate Constant 1.0E-2
Reaction Order 0.5
Surface Pump Rate 5 bbl/min
HCl Concentration 15%
Volume of HCl 5,000 gal
Total Treatment Time 23.8 min
Normalized Pump Rate 0.1 bbl/min/ft
Normalized Volume 100 gal/ft
Treatment Design
Table 2Parametric Study Information
Formation Characteristics
Formation Data Well No. 1 Well No. 2
Temperature 150F 150F
Perforated thickness 102 ft 66 ft
Permeability 100 md 5 md
Porosity 0.21 0.18
Reactivity (assumed) High High
Treatment Data
Plain 15% HCl 51 gal/ft 73 gal/ft
In-situ crosslinked (5% HCl) 37 gal/ft 51 gal/ft
Total acid volume 88 gal/ft 124 gal/ft
Pumping rate 0.1 to 0.4 bbl/min/ft 0.1 to 0.6 bbl/min/ft
Model Calibration
Average stimulation depth 6.8 ft 10.2 ft
Accessible porosity 8.7% 5.5%
Pore volumes 0.41 0.31
Post-Acid Skin -3.3 -3.7
Table 3Formation and Treatment Data for Calibration of New Theory
SPE 54719 9 R. GDANSKI
Fig. 1Radial symmetry of water wormholing in plaster casts (Daccord).
Fig. 2Side view of wormhole penetration for general example. Fig. 3Top view of wormhole penetration for general example.
om00237
om000239
om000238
10 SPE 54719 A FUNDAMENTALLY NEW MODEL OF ACID WORMHOLING IN CARBONATES
Fig. 4Wormhole radius profile for general example. Fig. 5Acid spending profile for general example.
Fig. 6Nominal 1/Da at wormhole tip for general example. Fig. 7Reynolds number at wormhole entrance for general
example.
Fig. 9Effect of reaction rate on wormhole development.
Fig. 10Effect of pumping rate on wormhole development.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0 20 40 60 80 100
Length (in.)
R
a
d
i
u
s


(
i
n
.
)
0
2
4
6
8
10
12
14
16
0 20 40 60 80 100
Length (in.)
H
C
l


(
%
)
0.1
1
10
0 5 10 15 20 25
Time (min)
1

/

D
a
0
20,000
40,000
60,000
80,000
100,000
0 5 10 15 20 25
Time (min)
R
e
y
n
o
l
d
s

N
u
m
b
e
r
0.0
0.2
0.4
0.6
0.8
1.0
0 50 100 150 200
Wormhole Length (in.)
W
o
r
m
h
o
l
e

R
a
d
i
u
s


(
i
n
.
)
2%
5%
10%
25%
50%
Fig. 8Effect of porosity on wormhole development.
0.0
0.2
0.4
0.6
0.8
1.0
0 20 40 60 80 100
Wormhole Length (in.)
W
o
r
m
h
o
l
e

R
a
d
i
u
s


(
i
n
.
)
1.0E-2
3.0E-3
1.0E-3
3.0E-4
1.0E-4
3.0E-5
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0 20 40 60 80 100
Wormhole Length (in.)
W
o
r
m
h
o
l
e

R
a
d
i
u
s


(
i
n
.
)
0.05
0.5
5.0
50
500
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0 50 100 150 200 250
Wormhole Length (in.)
W
o
r
m
h
o
l
e

R
a
d
i
u
s


(
i
n
.
)
50000
25000
10000
5000
500
Fig. 11Effect of acid volume on wormhole development.

S-ar putea să vă placă și