You are on page 1of 13

SPE 63183

Modeling of the Acidizing Process in Naturally-Fractured Carbonates


C. Dong, D. Zhu and A. D. Hill, University of Texas at Austin

Copyright 2000, Society of Petroleum Engineers Inc.


This paper was prepared for presentation at the 2000 SPE Annual Technical Conference and
Exhibition held in Dallas, Texas, 14 October 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
When a naturally fractured carbonate formation is treated with
acid at pressures below the fracturing pressure, the acidizing
process will likely be different from either a matrix treatment
or an acid fracturing treatment. Our previous experimental
study shows three kinds of acid etching patterns after acidizing
naturally-fractured carbonates and also illustrates their
relationships with the fracture properties and acidizing
conditions. Based on the experimental observations, a
mathematical model of acidizing naturally-fractured
carbonates has been developed. The model includes bulk
solution transport, acid transport and reaction, and the change
of fracture width by acid dissolution. A new approach was
used to treat leakoff acid and acid-rock reactions at the
fracture walls. The new acidizing model for numerically
generated rough-surfaced fractures predicts the same kinds of
acid etching patterns and the same relationships between acid
etching patterns and the fracture width, surface roughness and
leakoff rate as observed in experiments.
Introduction
Matrix acidizing is a stimulation method to improve the well
productivity by pumping acid at a pressure lower than
formation fracture pressure, and usually creates wormholes in
an un-fractured carbonate. Acid fracturing is another method
to improve the well productivity in carbonate reservoirs, in
which acid flows through the relatively wide hydraulic
fracture and etches the fracture walls. When a naturally
fractured carbonate formation is treated with acid at pressures
below the formation fracture pressure, the acidizing process is
neither like matrix acidizing nor like acid fracturing. Our
previous experimental studies1 show that in acidizing
naturally-fractured carbonates with a fracture width on the

order of 10-3 cm, most acid flows through and reacts with rock
inside the fracture, not the matrix. When the fracture width is
smaller than 210-3 cm, wormholes are created along the
fracture surface from the inlet to the outlet, which is like
matrix acidizing. When the fracture width is between 310-3
cm and 810-3 cm, a channel which is broad near the inlet and
narrower farther away from the inlet is created along the
fracture surface. When the fracture width is larger than 110-2
cm, acid etches most of the fracture surface like in acid
fracturing. When the fracture surface becomes rougher,
wormholes are more easily created. Leakoff has little effect on
the etching pattern along the fracture surface but creates
wormholes perpendicular to the fracture surfaces.
We developed a mathematical model to describe the
acidizing process in naturally-fractured carbonates. The model
is based on mass conservation for the acid solution, acid
transport and the change of fracture width by acid dissolution.
Pressure, fracture width and acid concentration as functions of
position and time can be predicted by the model. The
simulation results compared well with the experimental
observations.
Model Development
We initially formulated the model for the laboratory scale in
order to make direct comparisons with our experimental
results. The model domain is a block of carbonate core sample
with a length, l, a thickness, 2wm and a height, h, as shown in
Fig. 1. A single fracture is placed in the middle and through
the entire core sample. Acid is introduced from one side, flows
along the fracture, and flows out of the fracture from the
opposite side. The leakoff acid penetrates the matrix
perpendicular to the fracture. The coordinate system is defined
such that the x direction is the effluent acid flow direction,
which is aligned with the length of the fracture, the z direction
is aligned with the height of the fracture and the y direction is
perpendicular to the fracture surface. The fracture plane is the
x-z plane and the leakoff is in the y direction. The control
volume is a parallelepiped with dimensions of x, z and b,
where b is the fracture width at the point (x, z) (Fig. 2).
Mass Conservation. A mass balance for acid solution inside
the fracture is

C. Dong, D. Zhu, A. D. Hill

(v x b ) (v z b )
b

2v l =
x
z
t

(1)

where vx and vz are the average velocities in the fracture at


the point (x, z) in the x and z directions respectively, vl is the
leakoff velocity and b is the fracture width at the point (x, z).
If laminar flow is assumed, then the permeability of the
fracture at a certain position is b 2 12 , and the average
velocities in the fracture width vx and vz can be calculated by
vx =

b 2 p
12 x

(2)

vz =

b 2 p
12 z

(3)

where is acid viscosity and p is the pressure at the point


(x, z).
The leakoff velocity vl is an important parameter in Eq. 1. In
our experiments, the leakoff acid penetrates the rock matrix in
the y direction, and depends on the pressure drop applied
across the core. Thus, in the laboratory scale, the leakoff
velocity is proportional to the pressure difference at the point
(x, z) in the y direction, which is calculated by Darcys law as
vl =

k p pe
wm

1 3 p
3 p
b
+
b

x x 12 y z
p pe b
=
wm
t

(5)

Acid transport. Acid transport in the fracture is given by

( ) ( )

C
y

y=

b
2

= k g C Cw

(7)

where C is the acid concentration, D is acid diffusion


coefficient and Cw is the acid concentration at the fracture
walls. kg has the same units, m/s, as velocity, and accounts for
the rate of acid flux to the fracture walls by diffusion. When
hydrochloric acid is used to acidize carbonates as in our
experiments, the reaction rate is so high that Cw can be treated
as zero.
The first two terms of Eq. 6 are the acid transport caused by
the acid flow in the x and z direction. The last two terms are
the acid transport out of the fracture in the y direction by
leakoff and diffusion respectively.
Substituting Eqs. 2 - 4 into Eq. 6 yields
1 3 p
1 3 p

Cb
+
Cb
z
x 12 z
12 x

( )

k p pe
Cb
2Ck g =
2C
wm
t

(8)

Equation 8 represents acid transport during the acidizing


process.

Equation 5 represents transport of the acid solution.

where C is the average acid concentration across the


fracture width at the point (x, y) and kg is the apparent masstransfer coefficient, defined by Roberts and Guin2 as

(4)

where k is the permeability of the core matrix, pe is the back


pressure and wm is the half thickness of the core.
According to Eq. 4, the leakoff velocity changes with
position and time. Since the pressure near the inlet is higher
than farther away from the inlet, the leakoff rate near the inlet
should be higher than other areas. Thus, wormholes in the
leakoff direction are most probably created near the inlet,
which is consistent with the experimental observations.
Combining Eqs. 1 - 4 yields
1
12
k
2

SPE 63183

( )

Cv x b Cv z b
Cb

2Cvl 2Ck g =
x
z
t

(6)

Change of the fracture width. The conventional way to treat


the acid transport in the y direction is shown in Fig. 3. Acid is
transported to the fracture walls by leakoff and diffusion, and
before leaving the fracture, all leakoff acid and diffusion acid
contacts and reacts with the fracture walls. No wormhole is
created perpendicular to the fracture. If this is true, the etching
of the fracture wall increases as leakoff rate increases.
However the experimental results showed that there was no
significant difference in acid etching between with and
without leakoff. Furthermore, almost all leakoff acid flows
through wormholes not through the fracture walls uniformly,
which implies that most leakoff acid flows out of the fracture
without reacting with the fracture walls. Thus, a new approach
was used in the model as shown in Fig. 4.
Leakoff acid is highly unevenly distributed. Most leakoff
acid flows through wormholes perpendicular to the fracture,
and does not react with the fracture walls, as shown in Fig. 4.
Only a small part of the leakoff acid reacts with the fracture
walls before leaving the fracture. Since diffusion acid does not
create wormholes, all acid transported by diffusion reacts with
the formation at the fracture walls. With these assumptions,
the change of the fracture width is represented by
2k g C + 2 Cvl =

b (1 )

(9)

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

where is the rock density, is the rock matrix porosity,


is the dissolving power of acid and is the fraction of leakoff
acid that reacts with rock at the fracture walls. The two terms
on the left hand side of Eq. 9 represent the rate of acid
transported to the fracture walls and reacted with rock at the
fracture walls by diffusion and by leakoff. All the acid
transported by diffusion reacts with rock on the fracture
surface, while only a small part of the leakoff acid reacts at the
fracture surface. For our experimental conditions, << 1.
Substituting Eq. 4 into Eq. 9 and rearranging yields
b
k p pe
2
C + 2k g C =
(1 ) wm
t

(10)

Equation 10 represents the change of the fracture width by


acid dissolution.
Equations 5, 8 and 10 form the theoretical basis of the
model. Three unknowns p, b and C change with both position
and time, and can be calculated by solving the system of
equations.
Initial and boundary conditions. In our laboratory
experiments, before acidizing, inside the fracture the pressures
are the back pressure and acid concentrations are zero. When
acidizing begins, acid is introduced into the fracture with a
constant injection rate from the inlet side at x = 0. There is no
flow at the top and bottom sides where z = 0 and z = h. At the
outlet side where x = l, the pressure is the back pressure. The
acid concentration at the inlet side is Ci. Eqs. 11 - 18 represent
the initial and boundary conditions for our laboratory
experiments.
p ( x, z ) = 0 x, z , t = 0

(11)

C ( x, z ) = 0 x , z , t = 0

(12)

b( x, z ) = f ( x, z ) at t = 0

(13)

b 3 p

0 12 x x =0 dz = q
p (l , z ) = p e
p
z

z =0

p
z

z =h

t >0

z, t > 0

(14)
(15)

= 0 x, t > 0

(16)

= 0 x, t > 0

(17)

C (0, z ) = Ci z, t > 0

(18)

where q is the injection rate.


Generation of Rough-Surfaced Fractures
To solve the system of equations, we numerically generate the
rough-surfaced fracture with unevenly distributed fracture
widths as the one of the initial conditions. First, rough fracture
surfaces are generated by a method called the random mid-

point displacement method whose algorithm is based on the


fractional Brownian motions proposed by Saupe3. Two
parameters, fractal dimension and initial standard deviation of
the rough surface are needed to generate a rough surface. Each
rough surface is divided into rectangular elemental areas.
Within each elemental area, the elevation over the surface
base is the same, and the surface roughness is caused by
different elevations of different elemental areas. Figs. 5-7
show the numerically generated rough surfaces with the same
standard deviation of 0.01 cm but different fractal dimensions
ranging from 2.0 to 2.5.
Two surfaces are combined together to form a roughsurfaced fracture with a certain hydraulic width. In most cases,
the hydraulic width is not equal to the real average fracture
width. Brown4 analyzed the fluid flow in rough-surfaced
fractures and found that the relationship between the hydraulic
width and the real average fracture width can be expressed by
3

bh

b =
av

1
b
1 + 0.6 av
ap

1.20

(19)

where bh is the hydraulic width, bav is the real average


fracture width and ap is the standard deviation of the fracture
width.
We use the method provided by Amadei and Illangasekare5
to generate rough-surfaced fractures. Two rough surfaces with
the same fractal dimensions and initial standard deviations but
different random surfaces are placed parallel to each other and
the distance between the two surfaces is set to the hydraulic
width. The fracture width (void space) can be calculated from
the difference in topographic elevation between the two
surfaces. At contact points, any overlap of the surface is
removed and the local width is set to the value corresponding
to the permeability of the rock matrix without fractures. The
real average fracture width bav and the standard deviation of
the fracture width are calculated. By checking the relation
between bav and bh with Eq. 19 and adjusting the distance
between the two surfaces, the initial fracture width distribution
with a certain surface roughness and a hydraulic width is
obtained. This fracture width distribution is input into the
simulation as one of the initial conditions.
Simulation
Results of Eqs. 5, 8, and 10 were solved numerically to
simulate the acidizing process in naturally-fractured
carbonates.
In the simulations, the hydraulic widths were in the range of
1.7 10-3 cm to 1.2 10-2 cm, which were the same as in the
experiments. We measured the fracture surface roughness
from the core samples and used these values in the
simulations. In both the experiments and the simulations, the
initial leakoff rate was controlled by the rock matrix
permeability and the back pressure. All other parameters such

C. Dong, D. Zhu, A. D. Hill

as the rock properties and the acidizing conditions were the


same as used in the experiments. The input parameters used in
the simulations are listed in Table 1.
For each simulation, the fracture was numerically generated
with a different set of random numbers. The dimension of the
fracture was 6.56.5 cm, and the grid blocks were 6565. The
fracture width distribution before and after acidizing were
plotted. Other data like pressure, flow rate and acid
concentration were also recorded, but here only the fracture
width data are presented.
Case 1. Base case. The hydraulic fracture width before
acidizing was 6.4 10-3 cm. The standard deviation and the
fractal dimension of the fracture surface were 1.0 10-2 cm
and 2.0 respectively. The initial leakoff rate was 0%. Fig. 8
shows the initial fracture width distribution before acidizing,
illustrating the random distribution of the fracture width. In
Fig. 8 and all the following pictures, the bottom side is the
acid inlet side and the topside is the outlet side.
The simulated width distribution after acidizing is shown in
Fig. 9. A channel was created, which was broad and deep near
the inlet and narrow and shallow near the outlet. The depth of
the channel is around 0.3 cm and the surface of the channel
changed because of the acid etching. Other fracture areas
beside the channel had little change.
This simulation result is very close to the experimental
result with the same acidizing conditions, as shown Fig. 10. In
both the simulation and the experiments, we observed the
same kind of etching patterns with the same characteristics.
Case 2. Increased fracture width. In case 2, the fracture
width before acidizing was increased to 1.2 10-2 cm, and all
other parameters were similar to case 1. The fracture before
acidizing contained more large fracture widths, which were
distributed unevenly as shown in Fig. 11.
After acidizing, a very broad channel was predicted, and the
entire inlet side was etched by acid as shown in Fig. 12. As
acid flowed away from the inlet, the channel converged and
became narrower at the outlet, but the average channel breadth
was larger than in case 1. Other areas beside the channel had
little change. In case 2, as the fracture width increased, the
channel became broader, and the initial fracture width
distribution had less effect on the final fracture width
distribution. Compared with the experimental result with the
same conditions shown in Fig. 13, the simulator predicted the
same kind of etching pattern as the fracture width increased.
Case 3. Decreased fracture width. In case 3, the fracture
width was decreased to 1.7 10-3 cm with all other parameters
the same as case 1. The fracture before acidizing contained
few large fracture widths, and the contact area became much
larger than before as shown in Fig. 14. Large fracture widths
were concentrated in the top-left and the lower-center parts of
the fracture.
Figure 15 shows the simulation result of case 3. A narrow
channel was created, which became gradually narrower farther

SPE 63183

away from the inlet, and finally split to a channel and a narrow
wormhole. The average channel breadth was smaller than in
case 1, and the channel was along the initial large fracture
widths before acidizing. In this case, when the fracture width
decreased, the channel became narrower and the initial
fracture width distribution before acidizing had more effect on
the final fracture width distribution, very similar to the
experimental results shown in Fig. 16.
The simulation results of cases 1, 2 and 3 predict the same
relationship between the acid etching pattern and the initial
fracture width as formed in the experiments.
Case 4. Increased leakoff. In case 4, the leakoff rate was
increased to 19.2%, and the other parameters were the same as
case 1. The initial fracture width distribution is shown in Fig.
17, which is like case 1.
After acidizing, a channel was created as shown in Fig. 18.
The channel depth was about 0.3 cm. The depth and the
average channel breadth were close to both the experimental
result shown in Fig. 19 and case 1. Thus in both simulation
and experiments we observed that increasing the leakoff rate
has little effect on acid etching patterns.
Case 5. Increased standard deviation of fracture surface.
In case 5, we increased the fracture surface roughness by
increasing the standard deviation of the fracture surfaces to 3
10-3 cm and keeping the other parameters the same as in case
1. The fracture contained larger contact area and the fracture
width distribution became more non-uniform as shown in Fig.
20.
After acidizing, a y shaped wormhole was created as
shown in Fig. 21. The average wormhole breadth was smaller
than in case 1. Other areas besides the wormhole had little
change. In this case, as the standard deviation of the fracture
surfaces increased, the channel became narrower and
wormholes were easily created. This was very close to the
experimental result shown in Fig. 22.
Case 6. Increased fractal dimension. In case 6, the fractal
dimension of fracture surfaces was increased to 2.5, and other
parameters were unchanged. The initial fracture width
distribution is shown in Fig. 23. The contact area became
much larger, and the fracture width distribution became more
non-uniform.
After acidizing, a narrow channel was created as shown in
Fig. 24. The channel followed the center part of the fracture
which contained the largest fracture width before acidizing.
The breadth of the channel did not change much and was
smaller than in case 1. Other areas besides the channel had
little change. In this simulation, as the fractal dimension
increased, the channel became narrower.
Combining the simulation results of cases 5 and 6, when the
fracture surface roughness increased, the channel became
narrower and a wormhole was more easily created. Also the
initial fracture width distribution before acidizing had more
effect on the final fracture width distribution. Thus in both

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

simulations and experiments, we observed the same


relationship between the channel breadth and the fracture
surface roughness.
CONCLUSIONS
A mathematical model has been developed to describe the
acidizing process in naturally-fractured carbonates. The model
is base on mass conservation of the acid solution, acid
transport and the change of fracture width by acid dissolution.
A new approach is used to treat leakoff acid and acid-rock
reactions at the fracture walls. Simulations of acidizing
numerically-generated, random-rough-walled fractures predict
channels, which are broad and deep near the inlet and narrow
and shallow farther away from the inlet. As the fracture width
decreases, or as the surface roughness increases, the average
breadth of the channels decreases and wormholes are more
easily created. When the initial fracture width increases, the
channel becomes broader and less affected by the initial
fracture width distribution. Leakoff rate has little effect on
etching patterns. In both experiments and the simulations, the
same relationships between acid etching patterns and fracture
width, surface roughness and leakoff rate are observed,
indicating that the description of acid transport and reaction
and the method of numerically generating rough-surfaced
fractures are correct.
NOMENCLATURE
b = fracture width (cm)
bav = average fracture width (cm)
bh = hydraulic fracture width (cm)
C = acid concentration (wt.%)
C = average acid concentration across the fracture
width at the point (x, y) (wt.%)
Cw = acid concentration at the fracture walls (wt.%)
D = diffusion coefficient (m2/s)
h = fracture height (cm)
k = rock matrix permeability (md)
kg = mass transport coefficient (m/s)
L = fracture length (cm)
p = pressure inside the fracture (psi)
pe = back pressure (psi)
q = acid injection rate (ml/min)
t = time (s)
vl = fluid leakoff velocity (cm/s)
vx = average velocity in the x direction across the
fracture width (cm/s)
vy = average velocity in the y direction across the
fracture width (cm/s)
wm = half thickness of the core sample (cm)
= acid solving power, (kg rock/kg pure acid)
= rock matrix porosity
= acid viscosity (cp)
= rock density (g/cm3)
= fraction of leakoff acid that reacts with the
rock at the fracture walls
ap = standard deviation of the fracture width (cm)

REFERENCE
1. Dong, C., Zhu, D. and Hill, A. D.: Acid Etching Pattern
in Naturally-Fractured Formations, SPE 56531 presented
at the 1999 SPE Annual Technical Conference, Houston,
Texas, 36 October 1999.
2. Roberts, L. D. and Guin, J. A.: A New Method for
Predicting Acid Penetration Distance, SPEJ (Aug. 1975)
277-285.
3. Peitgen H. O. and Saupe D.: The Science of Fractal
Images, Springer, Berlin 1988.
4. Brown S.: Fluid Flow Through Rock Joints: the Effect of
Surface Roughness, J. Geophys. Res. 92, 1337-1347
(1987).
5. Amadei, B. and Illangasekare, T.: A Mathematical
Model for Flow and Solute Transport in NonHomogeneous Rock Fractures, Int. J. Rock Mech. Min.
Sci. & Geomech., vol. 29, November 1994.

C. Dong, D. Zhu, A. D. Hill

SPE 63183

Table 1. Input parameters for simulations

Simulation

Hydraulic fracture
width before acidizing
( 10-3 cm )

Fractal dimension
of the fracture
surface

Initial standard
deviation of the
fracture surface
( 10-2 cm )

Initial
leakoff rate

Case 1

6.4

2.0

1.0

0%

Case 2

12.0

2.0

1.0

0%

Case 3

1.7

2.0

1.0

0%

Case 4

6.7

2.0

1.0

20%

Case 5

6.0

2.5

3.0

0%

Case 6

6.9

2.0

1.0

0%

Fig. 1. Model domain: a carbonate core sample with a


single fracture

Fig. 2. Control volume in the fracture

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

Fig. 3. Conventional treatment of leakoff and


diffusion acid

Fig.5. Numerically generated rough


surface with fractal dimension 2.0

Fig. 4. New approach to treat leakoff and


diffusion acid

Fig. 6. Numerically generated rough


surface with fractal dimension 2.2

Fig. 7. Numerically generated rough


surface with fractal dimension 2.5

C. Dong, D. Zhu, A. D. Hill

Fig. 8. Initial fracture width distribution before


acidizing in case 1

Fig. 9 Final fracture width distribution after


acidizing in case 1

Fig. 10. Acid etching pattern from


experiments

SPE 63183

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

Fig. 11. Initial fracture width distribution before


acidizing in case 2

Fig. 12. Final fracture width distribution after


acidizing in case 2

Fig. 13. Acid etching pattern from


experiments

10

C. Dong, D. Zhu, A. D. Hill

Fig. 14. Initial fracture width distribution before


acidizing in case 3

Fig. 15. Final fracture width distribution after


acidizing in case 3

Fig. 16. Acid etching pattern


from experiments

SPE 63183

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

Fig. 17. Initial fracture width distribution before


acidizing in case 4

Fig. 18. Final fracture width distribution after


acidizing in case 4

Fig. 19. Acid etching pattern from


experiments

11

12

C. Dong, D. Zhu, A. D. Hill

Fig. 20. Initial fracture width distribution before


acidizing in case 5

Fig. 21. Final fracture width distribution after


acidizing in case 5

Fig. 22. Acid etching pattern from


experiments

SPE 63183

SPE 63183

MODELING OF THE ACIDIZING PROCESS IN NATURALLY-FRACTURED CARBONATES

Fig. 23. Initial fracture width distribution before


acidizing in case 6

Fig. 24. Final fracture width distribution after


acidizing in case 6

13

You might also like