You are on page 1of 25

Journal of Petroleum Science and Engineering 186 (2020) 106776

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: http://www.elsevier.com/locate/petrol

Carbonate acidizing: A mechanistic model for wormhole growth in linear


and radial flow
Mahmoud Ali, Murtaza Ziauddin *
Schlumberger, United States

A B S T R A C T

Highly conductive fluid flow paths, known as wormholes, are formed when acids are injected in carbonate reservoirs during a stimulation treatment. In matrix
acidizing treatments, these wormholes are desired whereas in acid fracturing treatments, they are undesirable because they reduce efficiency of the process. Two
main types of models are available in the literature for wormhole growth. Semi-empirical models are computationally fast; however, they do not scale well to core
dimensions or to radial flow. Numerical simulation models represent the physics of the individual processes well; however, they are computationally intensive and
impractical to use on a field scale.
To address the shortcomings of these models, a computationally fast mechanistic model is developed for wormhole growth. The model scales well with flow rate
and core dimensions. It hence allows scale-up from linear laboratory core flow experiments to field scenarios of radial flow. Accurate scaling with core dimension and
with injection rate in linear flow is a necessary precondition for a model to be considered for scale-up to radial flow where both cross-sectional area to flow and the
fluid Darcy velocity change as the wormholes grow. Previous computationally fast models did not satisfy this precondition and therefore are not suitable for scaleup
to radial flow.
The model is validated against more than 50 published and internal sets of linear core-flood experiments on limestone and dolomite cores treated with HCl and
emulsified acids. The validation dataset covers a wide range of acid concentrations, temperatures, rock types and core dimensions. The modeling equations extend to
radial flow based on the change in the domain area with wormhole progression. The radial model was tested against 10 published radial flow experiments and a
reasonable match was obtained. The model predictions were also tested against field results and again a good match was obtained.

model has two tuning parameters to match linear experiments and up­
scales to radial flow using the fractal dimension introduced by Daccord
1. Introduction et al. (1989). Buijse and Glasbergen (2005) introduced a simple model
(“BG model”) for designing acid treatments. The model was based on
Carbonate acidizing is a well stimulation technique in which reactive matching experimental data using a semi-empirical equation. The BG
fluids are injected either below (matrix) or above (acid fracturing) for­ model requires information about optimum conditions from linear lab­
mation fracturing pressure to improve production from the well. Acid­ oratory experiments to design field treatments. Tardy et al. (2007)
izing of carbonates has been extensively studied (Lund et al., 1973, conducted both linear and radial experiments, and the results of the
1975; Daccord et al., 1993; Wang et al., 1993; Fredd and Fogler, 1998, study demonstrated that the BG model underpredicts acid volumes
1999; Bazin, 2001; Panga et al., 2002, 2005; Buijse and Glasbergen, required in radial flow. Furui et al. (2012) conducted a comprehensive
2005; Taylor et al., 2006; McDuff et al., 2010; Dong et al., 2014; Zakaria study based on experiments, field data, and numerical simulations.
et al., 2015b; Maheshwari and Balakotaiah, 2013; Qiu et al., 2018; Panga et al. (2005) introduced the two-scale continuum (TSC) model as
Dong, 2018). Daccord et al. (1989) introduced a wormhole propagation a powerful tool to simulate acid propagation. Maheshwari and Balako­
model based primarily on linear and radial experiments with plaster and taiah (2013) used the TSC model to simulate wormhole growth in 3-D.
water. The model upscales radially based on the fractal dimension, Ali and Nasr-El-Din (2019) used the TSC model with a porosity distri­
which was found to be 1.6. Economides et al. (1994) claimed that bution obtained from X-ray computed tomography (CT) scanner to
Daccord’s model predicts very deep wormholes and followed a conser­ simulate laboratory experiments and predict acid performance under
vative approach that assumes that acid dissolves a constant fraction of field conditions. They concluded that the relation between laboratory
the rock. The dissolved fraction is determined based on linear experi­ experiments and acid propagation under field conditions is not linear.
ments. Gong and El-Rabba (1999) proposed a wormhole growth model However, the TSC model requires enormous processing power, and
that accounts for acid convection, reaction rate, and diffusion. The

* Corresponding author.
E-mail address: ziauddin@slb.com (M. Ziauddin).

https://doi.org/10.1016/j.petrol.2019.106776
Received 30 July 2019; Received in revised form 31 October 2019; Accepted 2 December 2019
Available online 6 December 2019
0920-4105/© 2019 Elsevier B.V. All rights reserved.
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Nomenclature nwh Number of wormholes


P Pressure, kg/m2
A Cross-sectional area at any time, m2 PV Pore volume
a Enhanced permeability zone flow coefficient, (m)2(1–n) PVBT Pore volume to breakthrough
Am Injection face matrix area, m2 qm Flow rate inside the matrix (enhanced permeability zone),
Ao Injection face area, m2 m3/s
av Interfacial area, m2/m3 qo Injection rate, m3/s
b Wormhole flow coefficient, (m/s) 1 qopt Optimum injection rate, m3/s
c Dimensionless acid concentration qwh Flow rate inside the wormhole, m3/s
CA Acid concentration at the tip of the wormhole, fraction R Universal gas constant, kcal/(k.mol)
CAo Acid concentration at the inlet of the wormhole, fraction r Dimensionless core radius
CT Computed tomography rc Core radius, m
c/c0 normalized tracer concentration rp Pore radius, m
D Fluid based diffusion coefficient, dimensionless rw Wellbore radius, m
Da Damkohler number, dimensionless rwh Wormhole radius, m
Dm Diffusion coefficient, m2/s rwh.opt Wormhole radius at the optimum injection rate, m
DmConway Diffusion coefficient reported by Conway et al. (1999), Sh∞ Asymptotic Sherwood number, dimensionless
m2/s t Total injection time, s
Ea Activation energy, kcal/mol T Temperature, Kelvin
f Flowing fraction, dimensionless TSC Two scale continuum
h Dimensionless thickness v Fluid velocity inside the wormhole, m/s
ho Payzone thickness, m vis Fluid kinematic viscosity, m2/s
K Matrix (enhanced zone) permeability, m2 vo Darcy velocity, m/s
keff Effective reaction constant, 1/s X Acid volumetric dissolving power, dimensionless
km Mass transfer coefficient, 1/s γ Dimension correction parameter, (m)2(1–n)
ko Mass transfer coefficient static constant, 1/m2 β Dimensionless wellbore radius
ks Reaction rate constant, 1/s λ Dimensionless length
kso Pre-exponential factor τ Dimensionless time
L Wormhole (core) length, m Φt Rock total porosity, fraction
L Characteristic length, 1 m ω Dimensionless velocity
M Characteristic time, 60 s μ Fluid dynamic viscosity, kg/(m.s)
n Enhanced permeability area factor

the tip is estimated by splitting the total injected fluid into a matrix and
wormhole flow component. The resulting model has an unprecedented
capability of matching many linear coreflood experiments for a wide
range of temperatures, acid concentrations, mineralogy, and core di­
mensions with a single set of model parameters. The model is extended
to radial flow by changing the cross-sectional flow area as the wormhole
propagates deeper in the reservoir. A key strength of the model is that it
uses the same parameters for linear and radial flow. Therefore, once
these parameters are determined from linear core flow tests, they can be
used directly in predicting behavior in radial flow. Furthermore, since
the model is a mechanistic model, the parameter values allow a deeper
understanding of the wormholing phenomenon and allows comparison
of various acid-rock systems on an equivalent basis.

2. Linear model

The acid efficiency curve along with different dissolution patterns for
a typical linear coreflood acidizing experiment is shown in Fig. 1. Note
Fig. 1. Acid efficiency curve with different dissolution patterns using 0.5 M HCl that an optimum injection rate exists, at which a minimum pore volume
at room temperature (after Fredd and Fogler, 1996). of acid is required to break through (Fig. 1c). Fig. 2 shows a higher
resolution image for a similar experiment for another core near the
routine field-scale simulations are impractical. optimum. At optimum injection rate, there is limited leak off and hence
The objective of the present study is to develop a simple mechanistic branching from the dominant wormhole channel is also limited. The
model for carbonate acidizing that is 1) scalable with core dimensions, diameter is constant along the channel and is a characteristic of the acid-
linear to radial, and rock type; 2) valid for limestone and dolomite and 3) rock system and temperature. For injection rates much lower than the
applicable to all commonly used acid systems. optimum (Fig. 1a), the base of the wormhole near the entrance is wider,
The model presented here is based on calculating the acid concen­ but quickly establishes a channel of constant diameter. There is negli­
tration and fluid velocity at the wormhole tip. The wormholes are gible leak off from the channel and most of the acid spends before it
modeled as tubular reactors. Acid spends by reaction with the walls of reaches the tip. If the flow rate is reduced even further, the length of the
the wormhole as it travels down the wormhole channel. Fluid velocity at wormhole base increases. Eventually, for flow rates approaching zero

2
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 2. CT scan image of wormhole structure in limestone core using 15% HCl at 150 � F (Zakaria et al., 2015b).

Fig. 3. Conceptual model of wormhole propagation in linear cores.

the wormhole disappears completely and the pore volume of acid to optimum injection rate. Matrix injection of acid at rates higher than 2
breakthrough (PVBT) is simply the volume of acid required to orders of magnitude above the optimum is often not possible, because
completely dissolve the core. For intermediate flow rates (Fig. 2b), the the pressure required to do so is typically above the fracturing pressure
dissolution pattern is between that of Fig. 2a&c. As the injection rate is of the rock. These high rates are also difficult to realize for matrix
increased above the optimum, the fraction of acid flowing through the acidizing treatments in the field and are also not recommended due to
enhanced permeability matrix around the wormhole increases. The fluid inefficient use of the acid and a higher horsepower requirement. Acid
velocity in the wormhole, therefore does not increase proportionally injection at extremely low rates (i.e. those approaching zero) are also
with the injection rate (Fig. 2d). The wormhole width increases seldom measured directly in the laboratory, because PVBT can be esti­
marginally above the optimum. Instead, extensive branching is mated much more simply by the dissolving power of the acid. Therefore,
observed. As the injection rate is increased further, the flow of acid in experimental studies generally focus on injection rates where worm­
the rock matrix becomes dominant and the distinction between the holes are observed. The objective of the present study is to develop a
wormhole and matrix disappears (Fig. 2e). numerically efficient mechanistic model for this region.
Note that dominant wormholes are observed at least up to 1 order of A conceptual model of wormhole growth is shown in Fig. 3. The acid
magnitude lower and up to 1 order of magnitude higher than the flows down the wormhole channel with almost constant velocity and

3
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

reacts with the wormhole walls as it travels from the inlet to the tip. The 0 1
acid that reaches the tip is consumed in the wormhole longitudinal qo ð1 ϕt Þ @ keffv lc
growth and in stimulating the matrix around the wormhole channel. It is PVBT ¼ e 1A (9)
A0 lc ϕt CAo Xkeff
assumed that the entire rock porosity is accessible to the acid. The mass
of rock (dmR) in kg that is dissolved for the wormhole to grow by a length Zakaria et al. (2015b) introduced the concept of a flowing fraction (f)
dl in m can be expressed as to account for rock type and heterogeneity; the flowing fraction is
defined as the injected pore volume of an inert fluid that corresponds to
dmR ¼ ðaw ð1 ϕt ÞρR þ am ð1 ϕt ÞρR xÞ dl (1)
the normalized tracer concentration at CC0 ¼ 0:5. The flowing fraction is a
where aw is the cross-sectional area of a wormhole in m2, am is the cross- measure of the rock porosity that participates in fluid flow. It is
sectional area of the enhanced matrix around a wormhole, ϕt is the total considered the pore volume to breakthrough of an inert fluid (brine) for
porosity of the rock in fraction, ρR is the rock density in kg/m3, and x is the core and serves to normalize the observed PVBT. The same concept
the fraction of rock dissolved in the enhanced matrix around the can be applied here to relate PVBTs of two rock types (where subscripts 1
wormhole (Fig. 3). and 2 represent parameter values for rock type 1 and 2, respectively):
For a clearly distinguishable wormhole from the matrix, x ≪ 1, Eq.
PVBT1 PVBT2
(1) can be simplified to ¼ (10)
f1 f2
dmR ¼ ðaw ð1 ϕt ÞρR Þ dl (2)
The model equations are made dimensionless by defining the
Mass of acid (dmA) in kg consumed in a time dt in seconds (fast re­ variables:
action), is given by keff L CA l t rc v
Da ¼ ​ ; c¼ ​ ; λ¼ ; τ¼ ; r¼ ; ω¼
dmA ¼ CA ρA vaw dt (3) v CAo L M L vo

where CA is the acid concentration at the tip in weight fraction, ρA is the where Da is Damko €hler number, λ is the dimensionless length, L is the
fluid density in kg/m3, and v is the velocity of the acid at the tip in m/s. characteristic length, τ is the dimensionless time, M is the characteristic
The dissolving power of acid (β), on a mass basis is defined as: time, r is the dimensionless radius, rc is the core radius in m, ω is the
dimensionless velocity, and vo is the Darcy velocity in m/s.
β¼
dmR ð1 ϕt ÞρR dl
¼ (4) The dimensionless form of the PVBT equation is
dmA CA ρA vdt
ð1 ϕt Þ �
PVBT ¼ eDaλ 1 (11)
rearranging the equation ϕt λCAo X ωDa

dl vCA X
¼ (5) 2.1. Estimation of effective reaction rate
dt ð1 ϕt Þ

where X ¼ βρρA is defined as volumetric dissolving power of the acid. A The dissolution of the carbonate rock by acid is controlled by the rate
R
of Hþ transfer from the bulk acid to the wormhole wall as well as the
similar development of Eqs. (1)–(5) can be found in Hung et al. (1989)
surface reaction rate at the wall. Assuming local equilibrium between
and Furui et al. (2012). The spending of acid on the wormhole walls is
these two processes and assuming a first-order surface reaction
illustrated in Fig. 3. For a first-order irreversible reaction, the mass
balance on the acid yields ks Cs ¼ km ðCA Cs Þ (12)
dCA keff CA where ks is the surface reaction rate constant, Cs is the acid concentra­
¼ (6)
dl v tion at the rock surface in weight fraction, and km is the mass transfer
where keff is the effective reaction rate constant in 1/seconds Aris coefficient in 1/s. Solving for Cs .
(1999). Two mechanisms affect the fluid velocity in the wormhole km CA
Cs ¼ (13)
channel. The generation of carbon dioxide by the reaction increases the km þ ks
fluid volume and hence speeds up the fluid, whereas the fluid leakoff
Substituting back in Eq. (13)
from the walls slows it down. The net change in fluid velocity is assumed
to be minor. To simplify the model equations, the fluid velocity in the km ks
k s Cs ¼ CA ¼ keff CA (14)
wormhole is assumed to be constant for linear flow. Integrating and km þ ks
solving for CA
keff l
where keff ¼ kkmmþk
ks
s
. The effective reaction rate constant (keff ) is an overall
CA ¼ CAo e v (7) reaction rate constant and accounts for both surface reaction and mass
transfer rates in 1/s.
where CAo is the acid concentration at the inlet in weight fraction, and l
is the length of the wormhole in m. Substituting for CA from Eq. (7) in
2.2. Estimation of mass transfer constant (km)
Eq. (5), integrating and solving for time, t in seconds for the wormhole to
propagate a distance l
Arashi et al., (1982) developed a mass transfer correlation for a
0 1
monolith catalyst:
ð1 ϕt Þ @ keffv l
t¼ e 1A (8) � �1:36
CAo Xkeff 8km bA 16vbA
¼ 6:0 þ 0:0006 (15)
av Dm l*υis
Pore volume of acid to breakthrough (PVBT) is defined as AqoolctΦb t , and
can be calculated from Eq. (8) by using t ¼ tb and l ¼ lc (where tb is the where bA is the distance between the plates in m and υis is the kinematic
breakthrough time in seconds, and lc is the core length in m): viscosity in m2/s. Application of the above correlation to a wormhole
(assuming bA is approximately equal to wormhole diameter and l equals
the wormhole length) shows the effect of the convection term is minor
and can be neglected for most carbonate acidizing conditions.

4
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 4. Experimental diffusion data (dots) and generated correlation from Eq. (7) (line).

Accordingly, the mass transfer in the present study is approximated as � �


Ea
km � k0 Dm (16) RT
ks ¼ av kso e (19)
3av
where k0 � 4b A
. Wormholes are more tortuous, and their walls have where kso is the pre-exponential factor for the reaction in m/s, R is the
higher surface roughness compared to channels in a monolith catalyst. universal gas constant in Kcal/K.mol, Ea is the activation energy in Kcal/
The mass transfer is expected to be significantly faster in a wormhole, mol, and av is the interstitial area available for reaction per unit volume
3av
and hence the value of 4b A
should be considered as a lower bound for the of the medium in 1/m. For limestone-HCl system, the surface reaction
k0. Conway et al. (1999) developed a correlation for Dm between 294 K rate is orders of magnitude higher than the mass transfer rate (i.e., the
(70 � F) to 339 K (150 � F). Eq. (17) shows the Conway et al. correlation at mass transfer is the rate controlling step) for almost all practical acid­
initial condition of unspent acid (the original correlation accounts for izing scenarios. Consequently, keff ​ e ​ km is an excellent approximation
calcium and magnesium cations in solution as well): and the precise value of ks is not needed. For dolomites, the surface
� � reaction is rate-limiting at low temperatures, and ks must be included in
1621:411
DmConway ¼ exp þ 1:326CAo D (17) calculating the effective reaction rate.
T Lund et al. (1973) measured the surface reaction rate for the
HCl-dolomite reaction. They reported kso ¼ 9:4 � 1010 m/s, and Ea ¼
where T is the temperature in K, and D ¼ 14.205, 14.68, and 17.2
22.5 kcal/mol; these values are used in the present study. Furthermore,
for straight, gelled, and emulsified acid, respectively. Unfortunately, the
it was assumed that av is constant and equal to 220m-1.
Conway et al. correlation does not extend well to temperatures above
339 K (150 � F), especially for an HCl-limestone system. Therefore, in the
present study, their correlation was modified by refitting with higher 2.4. Fluid velocity inside a wormhole
temperature data (up to 416 K (290 � F)) from Fogler et al. (1973),
Roberts and Guin (1975), and Kotb et al., (2018) in addition to the An expression for the velocity inside the wormhole was developed
Conway et al. data. The modified expression for diffusion coefficient for based on the conceptual model shown in Fig. 2:
the HCl-limestone system used in the present study is q0 ¼ qm þ qwh (20)
� �
2270
Dm ¼ exp þ 1:326CAo 12:11 (18) where q0 is the total flow in the core in m3/s, qwh is the total flow in all
T
the wormholes in m3/s, and qm is the total flow in the matrix (including
The Conway et al. correlation and the modified correlation are stimulated matrix around the wormhole as well as the unstimulated
compared in Fig. 4. Higher temperature data for emulsified acid- matrix) in m3/s. qwh can be expressed in terms of average fluid velocity
limestone, HCl-dolomite, and emulsified acid-dolomite are not avail­ inside the wormhole (v):
able and, therefore, the Conway et al. correlation is used for these sys­
qwh ¼ vnwh πr2wh (21)
tems. The accuracy of the present model especially above 150 � F can be
improved for these systems when such data become available. Like Dong
where nwh is the number of the wormholes, and rwh is the average radius
et al., 2018 study, the diffusion coefficients in the present study are
of the wormhole in m. qwh can be expressed in terms of pressure drop
calculated at initial condition of unspent acid.
(ΔPwh) in Pa using the Hagen-Poiseuille equation:
2.3. Surface reaction rate nwh πr4wh ΔP
qwh ¼ ¼ vnwh πr2wh (22)
8μl
The reaction rate constant ðks Þ in 1/s can be expressed as
where μ is the fluid viscosity in Pa.s. Similarly, qm can be expressed in
terms of pressure drop (ΔPm ) in Pa using Darcy’s equation:

5
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 6. Inlet (top) and outlet core faces and wood’s metal casting of wormhole
in limestone core after acidizing. The blue line indicates the wormhole path
while the shaded area indicates the enhanced matrix zone (after Hoefner and
Fogler, 1989). (For interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of this article.)

Fig. 5. Wormhole propagation in radial flow. Top, actual CT scan image (after
McDuff et al., 2010). Bottom, a conceptual model.

Table 1
Values of the adjusted parameters for the studied acid-mineral systems. The
same parameter values are used for both linear and radial flow.
n a (m)2 b (s/m) ko (1/
(1–n)
m2 )

Limestone
HCl (without 0.65 2.68E- 17.3 2.43
CI) 04 Eþ07
HCl (with CI) 5.10E- 35.1 2.43
04 Eþ06
Emulsified 4.61E- 15.5 2.85
acid 04 Eþ07 Fig. 7. Core inlet face (top) and after acidizing X-ray CT scan (bottom). The
Dolomite T < 358.15 T>
blue line indicates the wormhole path while the shaded area indicates the
358.15
enhanced matrix zone (after Ali and Nasr-El-Din, 2019). (For interpretation of
HCl (without 0.75 8.28E- 1.65 Eþ12exp 16.5 1.09
CI) 03 (–0.09T) Eþ06 the references to colour in this figure legend, the reader is referred to the Web
HCl (with CI) 3.96E- 1.66 Eþ14exp 50.4 version of this article.)
03 (–0.081T)
Emulsified 3.31E- 1.2 Eþ14exp 36.4 3.00 KAm ΔP
acid 03 (–0.081T) Eþ06 qm ¼ (23)
μl

where K is the effective matrix permeability in m2 and Am is the effective


Table 2 matrix cross-sectional flow area in m2. Dividing Eq. (22) by Eq. (23) and
Radius of wormhole and permeability of the enhanced matrix using Eqs. (29) assuming the transverse pressure gradient is negligible ðΔPwh � ΔPm Þ
and (30). Values reported for 1.5 � 6 in. cores at 150 � F.
a (m)2(1–n)
qwh nwh πr4wh
b (s/m) rwh:opt (m) K (D) ¼ (24)
qm 8KAm
Limestone
HCl (without CI) 2.68E-04 17.3 8.8E-04 7700 Simulations with the TSC model confirm that the transverse pressure
HCl (with CI) 5.10E-04 35.1 7.5E-04 10,132 gradient is small (Maheshwari and Balakotaiah, 2013). Substituting
Emulsified acid 4.61E-04 15.5 4.9E-04 4154 back in Eq. (20) yields
Dolomite
� �
HCl (without CI) 8.28E-03 93.6 2.1E-03 78,020 8KAm
HCl (with CI) 3.96E-03 218.3 1.8E-03 27,358 q0 ¼ qwh 4
þ1 (25)
nwh πrwh
Emulsified acid 3.31E-03 91.7 4.2E-04 1216

6
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 8. (A) Effect of core diameter, (b) effect of core length from 4 in. to 8 in, (c) effect of core length from 6 in. to 20 in, and (d) comparison with a large core.

Substituting for qwh from Eq. (22) and solving for v yields terms of A as follows:
qo vo
v¼� � (26) v � n (29)
8KAm a AAo þ bvo
r2wh
þ nwh r2wh
π
where a and b are defined as
Assuming rwh � rwh:opt and nwh � qqopto
8Kγ
qo a¼ (30)
v¼ ! (27) r2wh:opt
8KAm
r2wh:opt
þ qqopto πr2wh:opt
π r2wh:opt
b¼ (31)
qopt
Dividing by Ao and simplifying yields

v ¼
vo
(28) which, in the dimensionless form, is
8KAm
r2wh:opt Ao
þ qvopto πr2wh:opt
1
ω¼� � (32)
�n
Am is the area of the enhanced permeability zone around the wormholes.
a
Ao
πL2 r2 þ bvo
It is intuitive that the ratio of the enhanced permeability area and the
cross-sectional area of the core decreases for larger cores, and it is al­ The expression of the velocity inside the wormhole completes the
ways <1. This enhanced permeability area is influenced by both the acid model. Differentiating Eq. (9) with respect to the injection rate
recipe and the rock type. In the present study, Am is approximated as 20 1 3
equal to γAn , where A is the cross-sectional area available for flow for dPVBT ð1 ϕt Þ keff lc aA n
k l keff lc
(33)
4@e v eff c
¼ 1A e v 5
given wormhole penetration in m2. The enhanced permeability area dqo A0 lc ϕt CAo Xkeff qo
factor, n captures the influence of the rock and acid type, while γ is
introduced to convert the units back to length squared. For linear cor­ The optimum injection rate, qopt is at dPVBT
dq
¼ 0. qopt can be found by
o
eflood experiments, A is constant and equals to Ao . This is because the solving Eq. (33) iteratively for dPVBT ¼ 0. However, a convenient explicit
dq
total cross-sectional area of wormholes is small compared to the core o

approximation to the solution is:


diameter for most injection rates. Therefore, Eq. (28) can be written in

7
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 9. (A) Effect of acid concentration, (b) CI, and (c) temperature.

Fig. 10. Effect of rock type is shown in a comparison between Zakaria et al. (2015b) experiments (dots) and model results (lines). (a) Data for Indiana Limestone,
Edwards Yellow, and Edwards White. (b) Data for Austin Chalk, Pink Desert, and Winterset. (For interpretation of the references to colour in this figure legend, the
reader is referred to the Web version of this article.)

8
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 11. Effect of back pressure. (a) Cheng et al., (2017) data and (b) Qiu et al., (2014) data.

Fig. 12. Effect of rock type is shown in a comparison between Zakaria et al. (2015a) experiments (dots) and model results (lines) at 150 � F.

1 þ 1:3 bq 2:6 b2q vo


qopt ¼ aq (34) v � n (35)
1 þ 2:7 bq a ð2πðrwAþlÞh oÞ
þ bvo
o

where aq ¼ aAn keff lc and. bq ¼ eð 1 bkeff lc Þ where ho is the pay-zone thickness in m. Note that fluid velocity inside
A calculation example for the PVBT and the optimum injection rate is the wormhole decreases as cross-sectional area available for flow in­
presented in Appendix A. creases. Eq. (35) can be written in the dimensionless form as follows:

3. Radial model ω¼�


1
� (36)
a
�n
2hπL2 ½β þ λ� þ bvo
McDuff et al. (2010) studied the effects of radial flow on carbonate
Ao

stimulation using large carbonate blocks. Fig. 5 shows a computed to­


mography (CT) image from one of their experiments. In the present where rw is the wellbore radius in m, h is the dimensionless thickness,
study, the linear model was extended to the radial flow by calculating defined as hLo , and β is the dimensionless wellbore radius, defined as rLw .
the cross-sectional area for acid flow as the wormhole propagates deeper The mass balance equation for the acid can be written in the
in the reservoir (Fig. 5). For wormhole of length l, cross-sectional flow dimensionless form for radial flow as
area A is equal to 2πðrw þ lÞho . Substituting in Eq. (29)
dc Lkeff 1
¼ dλ (37)
c vo ω

9
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 13. Comparison between Sayed et al. (2012) experiments (dots) and model results (line) at 300 � F.

Z λ
L eDaωα
τ¼ dλ (41)
Mvo CAo X 0 ω
It follows that the volume of acid required to reach a specified
wormhole penetration is given by
Z λ Daωα
LAo e
VA ¼ dλ (42)
CAo X 0 ω
Before moving to the validation section, it is worth noting that Eqs.
(11) and (42) were developed based on the following assumptions:

� The acid flows down the wormhole channel with a constant velocity
(v).
� The entire rock porosity is accessible to the acid.
� Local equilibrium between acid mass transfer rate and surface re­
action rate.
� First-order surface reaction.
� Interfacial area (av) is constant.
Fig. 14. Effect of temperature on HCl-dolomite system. Effective rate constant
(solid gray line) shows the transition of the process from reaction rate limited � The transverse pressure gradient between the wormholes and the
(dashed blue line) at low temperature to mass transfer limited (dashed orange matrix is negligible.
line) at high temperature. (For interpretation of the references to colour in this � The radius for dominant wormholes does not increase above the
figure legend, the reader is referred to the Web version of this article.) optimum injection rate.
� For injection rates higher than the optimum, the number of worm­
Integrating and solving for acid concentration holes is equal to the ratio of the injection rate and the optimum in­
� �n �n � jection rate.
Lkeff a 2hπL2 a 2hπL2 nþ1
ln c ¼ ðβ þ λÞnþ1 þ bvo λ β (38)
vo Ao ðn þ 1Þ Ao ðn þ 1Þ Relaxation of these assumptions may improve the prediction capa­
keff L bilities of the model but at the cost of increased complexity.
Recall that Da ¼ ​ v
and ω ¼ v
vo
. Rearranging and simplifying

c¼e Da ωα
(39) 4. Estimation of model parameters from linear coreflood data

" #
The parameters in the model were estimated by fitting the model to
að2hπ L2 Þn
where. α ¼ Ao ðnþ1Þ ððβ þ λÞnþ1 βnþ1 Þ þ bvo λ coreflood data reported in the literature (Table 1). Note that the same
parameter values are used for both the linear and radial versions of the
The equation of wormhole growth can be written in the dimen­ model. This allows for a straightforward scaleup from linear to radial
sionless form as flow, as parameters determined from linear flow measurements can be
dλ M used directly in the radial flow equations. The data covers HCl con­
¼ vo CAo X ωe Da ωα
(40) centrations between 0.5 wt% and 20 wt%, temperature from 72 � F to
dτ L
320 � F, core lengths from 4 in. to 20 in, and core diameters of 1 in.–7 in.
Integrating and rearranging
Both plain acid and emulsified acid are included. Data on plain acid

10
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 15. Effect of temperature. A comparison between Wang et al. (1993) experiments (dots) and model results (lines).

Fig. 16. Effect of temperature and acid concentration shown in a comparison between Ali and Nasr-El-Din (2019) experiments (dots) and model results (lines).

includes data with and without corrosion inhibitor (CI). Multiple rock mm and at the outlet it is 0.76 mm. Most of the wormhole channel ap­
types were included as well. Data on limestone acidizing are far more pears to be about 0.76 mm. The inlet radius is probably larger because it
abundant in the literature than on dolomite acidizing. Therefore, the includes the acid spent during the wormhole initiation phase, but it
confidence in the accuracy of the limestone parameters is higher than quickly converges to a radius of 0.76 mm. These values compare well
that for dolomite. The parameters in Table 1 are used in all the modeling with the estimated wormhole diameters of 0.88 mm and 0.75 mm in for
results presented in the paper. HCl/limestone system reported (Table 2). Fig. 7 shows the image of a
The accuracy of parameter values in Table 1 is assessed by calcu­ wormhole after acidizing a Silurian dolomite sample with 15 wt% HCl at
lating corresponding wormhole radii and enhanced matrix permeabil­ 150 Deg. F. The estimated wormhole radius from Fig. 7 is 2 mm, which
ities and comparing them to known values and trends (Table 2). compares well with computed wormhole diameter of 2.1 and 1.8 mm
The wormhole radii in Table 2 are calculated from Eq (30). The reported in Table 2. Note that the calculated wormhole radii for the
larger wormhole radii in dolomites compared to limestones are consis­ emulsified acid systems for both limestone and dolomite are smaller
tent with experimental measurements reported by Wang et al., Zakaria than that for the plain acid systems. This is consistent with the trend
et al. and Ali and Nasr-El-Din (Wang et al., 1993; Zakaria et al., 2015a, b; observed experimentally by the authors.
Ali and Nasr-El-Din, 2019). Fig. 6 show a Wood’s metal casting of a Direct measurements of the enhanced zone permeability are difficult,
wormhole obtained after acidizing Indiana limestone with plain HCl at and therefore a quantitative comparison is not possible here. The values
room temperature. The estimated wormhole radius at the inlet is 2.54 reported in Table 2 were calculated by first estimating the cross-

11
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

dolomites than for limestones (due to much more extensive wormhole


branching in dolomites). The high calculated permeability values sug­
gest presence of connected micro-wormholes in the enhanced matrix for
dolomites and only partially connected micro-wormholes for limestones.
The parameter n in the model is a characteristic of the rock and fluid
system and is comparable to half of the fractal dimension reported in
earlier studies (Daccord et al., 1989). The values of n reported here
compares well with the fractal dimension of 1.6 (n ~ 1.6/2 ¼ 0.8) re­
ported in these studies. A higher value of n is observed for dolomites
than for limestones which is consistent with more extensive wormhole
branching in dolomites than in limestones.

4.1. HCl-limestone

4.1.1. Effect of core dimension


The effect of core dimension on the required acid volume for
breakthrough is shown in Fig. 8a and b. Experimental data are depicted
by symbols and lines are used for modeling results. The core flow ex­
periments in Fig. 8a and b were conducted on Indiana limestone using
15 wt% HCl at room temperature (Dong et al., 2014). Fig. 8c shows the
comparison between experiments and model for Pink Desert limestone
for core lengths of 6 and 20 in. The experiments were conducted at room
temperature using 15 wt% HCl and 1.5-in. diameter cores. The data on
6-in. cores are from Zakaria et al. (2015b), and the data for 20-in. cores
are from Al-Ghamdi et al. (2011). Fig. 8d shows a comparison with data
from Seagraves et al. (2018). The core dimensions are 7 � 18 in.
The results show that the model scales well with core dimensions.
Accurate scaling of the model with core dimension and with injection
rate is a necessary precondition for the model to be considered for
Fig. 17. A comparison between CT scans of a high vug density dolomite core scaling up to radial flow where both cross-sectional area to flow and the
(Dong et al., 2018, left image) and a low vug density dolomite core (Ali and fluid Darcy velocity change as the wormholes grow. Previous models did
Nasr-El-Din, 2019, right image). not satisfy this precondition and therefore are not suitable for scaleup to
radial flow. For meaningful field predictions, previous models must be
sectional area of enhanced matrix zone from Wood’s metal casting of calibrated with radial flow data of comparable dimensions to the
wormhole or X-ray CT scans and then using Eq. (29) to calculate the wormhole penetrations observed in the field. Practically, this means that
permeability of the enhanced matrix. The calculated permeability values these models can only be calibrated with field data as large-scale lab
agree with the experimentally observed trends that (1) the permeability experiments or controlled injection experiments on carbonate outcrops
for the enhanced zone is much lower for emulsified acids than for plain are prohibitively expensive.
acids (possibly due to the lower leak off in the emulsified acid system),
and (2) the permeability of the enhanced matrix is much higher for

Fig. 18. Effect of temperature seen in a comparison between Dong et al. (2018) experiments (dots) and model results (lines).

12
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 19. A comparison between Sayed et al. (2013) experiment (dots) and model results (line) at 300 � F.

Fig. 20. A comparison between Zakaria et al. (2014) experiment (dots) and model results (line) at 230 � F.

4.1.2. Effect of acid concentration conducted on Indiana limestone at room temperature on 1.5-in. diam­
Fig. 9 shows the comparison between model predictions for 0.5 to 15 eter � 6-in. long cores with 15 wt% HCl acid. CI’s generally reduce the
wt% HCl at room temperature with experimental data from Wang et al. acid dissolution rate of carbonate by coating the rock surface and
(1993). Additionally, the model was compared with three sets of ex­ limiting access by acid. Therefore, as expected, the reaction rate without
periments with acid concentrations of 5, 15, and 20 wt% HCl at 150 � F a CI is about an order of magnitude higher than with corrosion inhibitor.
temperature using Indiana limestone and core dimensions of 1.5-in. The retarded reaction rate with the CI also results in a lower optimum
diameter � 6-in. long. A good match between the model and the ex­ injection rate. However, the results may not extend to scenarios where
periments was obtained for moderate to high acid concentrations. much larger wormhole lengths are expected, such as for long cores or to
However, for the very low acid concentration of 0.5 wt% HCl, the match field conditions. This is because the CI is expected to be adsorb rapidly
is not as good. The estimated model parameters in Table 1 do not ac­ near the entrance, and benefits may not be realized for deep wormholes.
count for change in acid activity with concentration. Most of the avail­ Further investigations into optimization of the fluid recipe with a CI or
able data are at high acid concentrations (i.e., typical acid other surface-active agents is beyond the scope of this study. However,
concentrations used in the field), and therefore the estimates are biased these effects can be easily accounted for in the model by adjusting the
towards these concentrations. reaction rate constant.

4.1.3. Effect of corrosion inhibitor 4.1.4. Effect of temperature


The effect of a CI on acid reaction rate was addressed by Rabie and Zakaria et al. (2015b) studied the effect of temperature on acid
Nasr El Din (2015). They studied three types of CI and found that sur­ performance as part of their study. Fig. 9c shows a comparison between
prisingly one type doubled the dissolution rate while others reduced the the Zakaria et al. (2015b) experiments and the present model results.
dissolution rate. Fig. 9b shows the effect of a CI on wormholing The effect of temperature was studied internally by conducting a
behavior. The data without CI are from Dong et al. (2014), and the data series of experiments at 200 � F, 275 � F, 300 � F, and 325 � F. A good match
with 1 vol% CI are from Zakaria et al. (2015b). Both experiments were was obtained between model and laboratory experiments.

13
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 21. Effect of reaction kinetics (temperature) on the optimum conditions and above optimum slope for HCl- and emulsified acid-limestone and dolomite systems.

14
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 22. A comparison between Tardy et al. (2007) radial experiments (dots) and radial model results (lines).

4.1.6. Effect of back-pressure


Qiu et al. (2014) found that the increase in the pressure from 1000
psi to 3000 psi results in a 50% decrease in the diffusion coefficient, due
to CO2 formed during the HCl-limestone reaction. Also, Cheng et al.
(2017) experiments indicate that a lower diffusion coefficient should be
expected at higher back-pressure. Fig. 11a and b shows a comparison
between the experimental data and the model for the back-pressure
data. In the calculations, the value of the diffusion coefficient at 3000
psi is half the value at 1000 psi (Qiu et al., 2014). The model captures the
change in the optimum injection rate but predicts higher PVBT values.
This can be attributed to the low CI concentration (0.2–0.3 vol%) used in
both studies. Also, an extensive study of the effect of pressure on the
diffusion coefficient may reveal lower values than reported by Qiu et al.
(2014).
Low values of PBVT are observed for back pressures below 1000 psi,
possibly due to reduced leakoff and increase in fluid volume due to
evolution of CO2. However, these low back pressures are not represen­
tative of most field treatments and were not considered in the model.

4.2. Emulsified acid-limestone

There is no standard recipe for emulsified acid, and seemingly small


variations in the recipe may lead to significant differences in perfor­
mance. For example, emulsified acids can be prepared at different ratios
of hydrocarbon and aqueous phases. The type of hydrocarbon (diesel,
dead oil, or kerosene) can be varied as well as the type of surfactant.
Furthermore, the preparation of an emulsified acid is a multistep pro­
cess, and a small variation in these steps can lead to large variations in
performance. For example, the speed and type of impeller used for
mixing hydrocarbon and aqueous phases as well as the mixing time
influence the droplet size and hence the reactivity of the system. Liter­
Fig. 23. A CT scan of an acidized limestone block. Only the blue part of the ature data were carefully selected to control for these parameters as
wellbore is open to the acid flow (after Aidagulov et al., 2018). (For interpre­ much as possible. However, as expected, the data from different labo­
tation of the references to colour in this figure legend, the reader is referred to ratories had a much larger scatter than data from the same laboratory.
the Web version of this article.) The literature data used in estimating the model parameters spanned
six rock types, temperatures between 150 � F and 300 � F, and acid con­
4.1.5. Effect of limestone rock type centrations between 15 and 28 wt% HCl. Zakaria et al. (2015a) con­
Zakaria et al. (2015b) studied the effect of rock type on acid per­ ducted acidizing experiments on six limestones rock types using
formance using six rock types. A comparison between the experimental emulsified acid. Fig. 12 shows a comparison between data from Zakaria
data and the model for Indiana limestone, Austin chalk, Edwards Yellow, et al. (2015a) and the model predictions. Fig. 13 shows a comparison
Pink Desert, Winterset, and Edwards White are shown in Fig. 10a and b. between data from Sayed et al., (2012) for low- and high-permeability
The current model assumes that all porosity is accessible (f ¼ 1). Indiana limestone at 300 � F. The model provided a reasonable match
Accordingly, the PVBT is calculated for the Indiana limestone (f ¼ 1) to most of the data except for 2 data points (marked with “?“). These 2
rock type and then calculated for the other rock types using Eq. (10). points do not follow the expected trend (no clear minimum PVBT) and

15
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 24. A comparison between Qiu et al. (2018) radial experiments (dots) and radial model results (lines).

Fig. 25. A comparison between Burton et al. (2018) radial experiments (dots) and radial model results (lines).

maybe are influenced by other factors. Additionally, the effect of con­ that need to be evaluated for dolomites is still larger than that for a
centration was tested with internal data sets, ranging from 15 wt% to 28 limestone system. Consequently, the accuracy of parameter values re­
wt% HCl and temperatures ranging from 200 � F to 280 � F. A reasonable ported in Table 1 for dolomites is also expected to be lower.
match with experimental data was obtained for these cases. Wang et al. (1993) studied the acidizing of San Andres dolomite
cores using 3.6 wt% HCl at room temperature, 122 � F, and 167 � F.
4.3. HCl and emulsified acid-dolomite Unlike wormhole behavior in limestones, the wormhole diameter in
dolomites changes significantly with temperature and cannot be
Evaluation of model parameters for HCl-dolomite and emulsified- assumed constant (Wang et al., 1993). Consequently, the a and b pa­
dolomite systems from core flow experiments is challenging because rameters in the model are defined as functions of temperature. Fig. 15
of a scarcity of literature data. For HCl, only three publications with shows a comparison with data from Wang et al. (1993). A flowing
enough experimental detail are available in the literature, and for fraction equal to 1 is used in calculating the model results. Wang et al.
emulsified acid, only two such publications exist (a total of 10 data sets did not report use of a CI in the acid recipe, and it is assumed that it was
are reported in these publications). Furthermore, as discussed earlier, not used in their experiments. The model parameters reported in Table 1
the effective reaction rate for dolomite is controlled by the surface re­ are with this assumption.
action at low temperature and by mass transfer at high temperature Ali and Nasr-El-Din (2019) conducted experiments on dolomite cores
(Fig. 14). As the wormholing mechanism shifts, the characteristic shape at 150 � F and 200 � F using 15 wt% and 20 wt% HCl, respectively. They
of the wormholes also changes. To reduce the number of model pa­ reported a flowing fraction 0.89 for this rock type. Fig. 16 shows a
rameters that need to be estimated from core flow studies, model pa­ comparison between Ali and Nasr-El-Din experiments and the model
rameters that could be determined independently of the core flow predictions.
studies, such as the surface reaction rate and diffusion coefficients, were Dong et al. (2018) studied dolomite acidizing at 122 � F, 185 � F, and
used as reported in the literature. However, the set of model parameters 260 � F using 15 wt% HCl. The studied range of temperatures covers the

16
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 26. Skin factor due to matrix acid treatments in carbonate reservoirs (Burton et al., 2018).

Fig. 27. Field-scale acid efficiency curves for multiple wormhole lengths. HCl-limestone case. 15 wt% HCl at 150 � F.

transition of the dolomite acidizing process from reaction-rate limited to 230 � F using 15 wt% HCl emulsified acid. A flowing fraction of 0.98 was
mass-transfer limited (Dong et al., 2018). The reported acid volumes are measured (Zakaria et al., 2014). Fig. 20 shows a comparison between
much lower than those reported by Ali and Nasr-El-Din (2019). Fig. 17 experiments and the model predictions.
compares the images from CT scans from the two studies. Note that the Additionally, HCl-dolomite model was tested against an internal set
core used by Dong et al. (2018) is much more heterogeneous (i.e., the of experiments at 325 � F using 1 � 6 in. cores, and a reasonable match
porosity is dominated by vugs) compared to the core used by Ali and was obtained.
Nasr-El-Din (2019). Therefore, the flowing fraction is expected to be
much lower. A flowing fraction of 0.35 was used in model calculations
for the Dong et al. (2018) cores. A comparison of experimental data and 4.4. Comparison of optimum injection rates for limestone and dolomites
model predictions is shown in Fig. 18.
Sayed et al. (2013) conducted coreflood experiments on Silurian It is instructive to compare these rock-acid systems on an eqiuvalent
dolomite cores at 300 � F using 15 wt% HCl emulsified acid. Fig. 19 basis using the model. Fig. 21 shows a comparison for core dimensions of
shows a comparison between experimental data from Sayed et al. (2013) 1.5 � 6 in. and 15% HCl. Different values of qopt were generated by
and the model predictions. No measurement for the flowing fraction was increasing the temperature from left to right in the figure. Note that:
reported by them. However, the rock type is like the one used by Ali and
Nasr-El-Din (2019), and therefore the same value of the flowing fraction � PVBTopt and dPVBT/dq (evaluated at q ¼ 3 qopt) are much higher for
(i.e., 0.89) is used in the model calculations. dolomite compared to limestone, especially at low temperature
Zakaria et al. (2014) conducted experiments on dolomite cores at (Fig. 21b and c). Therefore, a much larger acid volume is required for

17
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 28. Skin evolution with total acid volume injected at 0.8, 1.6, and 3.2 gal/(ft.min). HCl-limestone case, 15 wt% HCl at 150 � F.

Fig. 29. The change in acid optimum injection rate and acid volume with the wormhole length at 75, 150, and 300 � F. HCl-limestone system.

dolomites and the penalty from deviation from the optimum rate is 8 in. of the wellbore was open to acid flow. Fig. 23 shows the CT scan
also much higher. from the experiment. It is obvious that the wormholes traveled vertically
� For limestone the optimum PVBT increases monotonically with away from the openhole zone section. Therefore, the entire block height
temperature, while for dolomites there is an optimum temperature as (20 in.) is used in the model calculations. A comparison between the
well as an optimum injection rate (Fig. 21 b). experimental data and the model is presented in Fig. 24.
� Surprisingly, emulsified acids are much more effective in reducing Burton et al. (2018) performed several radial experiments on chalk at
the PVBT in dolomites than for limestones. 160 � F. The blocks had a thickness of 14 in. and a radius of 5 in.; a
wellbore of 0.5 in. radius was drilled in the middle of the block. Only 2
5. Validation against radial experiments in. of the wellbore was open to acid flow. Fig. 25 shows a comparison
between experiments and model results. A flowing fraction of 0.2 was
Tardy et al. (2007) performed two radial experiments at 75 � F and used in the model to account for the low acid volumes. The scatter in the
150 � F using 15 wt% HCl. The cylindrical cores were 2.25-in. thick with experimental data is attributed to (a) heterogeneous nature of chalk, (b)
a radius of 2.77 in.; a wellbore of 0.125-in. radius was drilled in the spherical flow observed in some experiments, (c) acid flow behind cas­
middle of the block. Fig. 22 presents the two experimental points from ing, and (d) the use of backpressure less than 1000 psi for some
Tardy et al. (2007), along with the model predicted volumes. experiments.
Qiu et al. (2018) conducted three radial experiments at 99 � F using
15 wt% HCl. The blocks had a thickness of 20 in. and a radius of 8 in.; a
wellbore of 0.5625-in. radius was drilled in the middle of the block. Only

18
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 30. Field-scale acid efficiency curves for multiple wormhole lengths. Emulsified acid-limestone case, 15 wt% HCl-emulsified acid at 150 � F.

Fig. 31. Skin evolution with total acid volume injected at 0.8, 1.6, and 3.2 gal/(ft.min). Emulsified acid-limestone case, 15 wt% HCl-emulsified acid at 150 � F.

6. Predicting field performance required to achieve deeper wormholes. This agrees with Ali and
Nasr-El-Din (2019) observations. Fig. 28 shows the development of the
Burton et al. (2018) collected historical carbonate acidizing data wormhole as a function of acid volume at three injection rates. The
from several hundred field treatments; the total volume of acid injected model predicts a skin factor of about 3.3 upon the injection of 100
per foot is in the range of 5–700 gal/ft, with a median of 75 gal/ft and an gal/ft, which is in a good agreement with field results (Fig. 26). The
average of 122 gal/ft. Post-treatment median skin values for these wells optimum conditions can be located using Fig. 27. The optimum injection
is 4 (Fig. 26). rate and acid volume as a function of wormhole lengths are shown at 75,
In this section, the model is used in predicting wormhole length and 150, and 300 � F in Fig. 29 the optimum injection rate and the wormhole
skin for radial flow. A 1-ft thick zone with a wellbore of 6-in. diameter is length can be predicted using the total acid volumes injected at a certain
studied. An open hole completion is assumed. The model parameters are temperature using Fig. 29. As an example, the injection of 100 gal/ft in a
the same as those determined for the corresponding linear core flow. 300 � F formation will result in a 4 ft wormhole at an injection rate of
about 0.5 gal/(ft.min). Fig. 29 is a unique acid design plot that can be
used to determine 1) the optimum injection rate and wormhole length
6.1. HCl-limestone for a given volume of acid, 2) the wormhole length and total volume of
acid for a given injection rate, and 3) the injection rate and volume for a
The model predictions for the linear coreflood experiments using 15 given wormhole length. The model can be easily used to run sensitivity
wt% HCl at 150 � F were presented in Fig. 9c. The radial acid efficiency analyses for the effect of the injection rate, total injected volumes, and
curves for different wormhole lengths (skin factor) are presented in pumping schedule. Similar plots can be generated for
Fig. 27. Not from Fig. 24 that higher optimum injection rates are

19
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 32. The change in acid optimum injection rate and acid volume with the wormhole length at 150, 220, and 300 � F, emulsified acid-limestone system.

Fig. 33. Field-scale acid efficiency curves for multiple wormhole lengths. HCl-dolomite case, 15 wt% HCl at 150 � F.

limestone-emulsified acid, dolomite-HCl, and dolomite-emulsified acid acid should be increasing with time following the optimum injection
systems. rates path to achieve best results. This observation agrees with recent
field treatments that follow the rate ramp-up design (Livescu et al.,
6.2. Emulsified acid-limestone 2018).

The model predictions for the linear coreflood Indiana limestone 6.3. HCl-dolomite
experiments using 15 wt% HCl at 150 � F were presented in Fig. 12. The
radial acid efficiency curves for different wormhole lengths (skin factor) The model predictions for the linear coreflood Silurian dolomite
are presented in Fig. 30. Emulsified acid requires higher optimum in­ experiments using 15 wt% HCl at 150 � F were presented in Fig. 16. The
jection rates to achieve deeper wormholes as well. Fig. 31 presents the radial acid efficiency curves for different wormhole lengths (skin factor)
development of the wormhole as a function of total acid volume injec­ are presented in Fig. 33. Like the HCl-limestone case, higher optimum
ted. A skin factor of 4 can be achieved by pumping 100 gal/ft. The injection rates are required to achieve deeper wormhole. The develop­
optimum injection rate and acid volume as a function of wormhole ment of the wormhole as a function of acid volume at three injection
lengths are shown at 150, 220, and 300 � F in Fig. 32. Following arrows rates is shown in Fig. 34. A skin of 2.5 is still achievable for the HCl-
in Fig. 32, the injection of 100 gal/ft in a 300 � F formation will result in a dolomite case. The developed wormholes are shorter than in the lime­
10 ft wormhole at an injection rate of about 0.8 gal/(ft-min). Comparing stone case, which is expected. The optimum conditions can be located
the HCl- and emulsified acid-limestone cases reveals that emulsified using Fig. 35. The optimum injection rate and acid volume as a function
acids will create longer wormholes, which agrees with laboratory and of wormhole length are shown at 75, 150, and 300 � F in Fig. 35.
field observations. Figs. 27 and 30 show that the injection rate of the Following the arrows in Fig. 35, the injection of 100 gal/ft in a 300 � F

20
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 34. Skin evolution with total acid volume injected at 0.8, 1.6, and 3.2 gal/(ft.min). HCl-dolomite case, 15 wt% HCl at 150 � F.

Fig. 35. The change in acid optimum injection rate and acid volume with the wormhole length at 75, 150, and 300� F. HCl-dolomite system.

formation will result in a 3.8 ft wormhole at an injection rate of about 300 � F in Fig. 38. Following the arrows in Fig. 38, the injection of 100
1.5 gal/(ft.min). Unlike the HCl-limestone system, the increase in for­ gal/ft in a 300 � F formation will result in an 8 ft wormhole at an injection
mation temperature for the HCl-dolomite system using the same volume rate of about 1.5 gal/(ft.min).
of acid is accompanied by deeper wormholes in the formation.
7. Conclusions

6.4. Emulsified acid-dolomite An experimentally-validated mechanistic model for carbonate acid­


izing is presented. The model accounts for acid consumption along the
The model predictions for the linear coreflood Silurian dolomite wormhole channel and the distribution of flow between the wormhole
experiments using 15 wt% HCl-emulsified acid at 300 � F were presented and the matrix. Inclusion of these two factors is critical in accurate
in Fig. 20. The radial acid efficiency curves for different wormhole scaleup with cores of different dimensions and from linear to radial flow.
lengths (skin factor) are presented in Fig. 36. Emulsified acid requires An analytical solution was obtained for the linear PVBT and was
higher optimum injection rates to achieve deeper wormhole as well. validated against 50 linear core flow tests covering a wide range of acid
Fig. 37 presents the development of the wormhole as a function of total type and concentration, injection rate, temperature and mineralogy. The
acid volume injected. A skin factor higher than 3.3 can be easily ability of the model to match a wide range of experiments with a single
achieved by pumping 100 gal/ft. The optimum injection rate and acid set of parameters validates the modeling assumptions. Furthermore, the
volume as a function of wormhole length are shown at 150, 220, and

21
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 36. Field-scale acid efficiency curves for multiple wormhole lengths. Emulsified acid -dolomite case, 15 wt% HCl-emulsified acid at 300 � F.

Fig. 37. Field-scale acid efficiency curves for multiple wormhole lengths. Emulsified acid-dolomite case, 15 wt% HCl-emulsified acid at 300 � F.

values of the model parameters were also independently estimated and � Deviation from the optimum injection rate for dolomite, especially at
found to be within an acceptable range. For radial flow, the model gave low temperature, results in much larger acid volume compared to a
reasonable predictions for 10 laboratory radial flow experiments and the limestone. PVBTopt and dPVBT/dq (evaluated at q ¼ 3 qopt) are much
treatment volumes predicted by the model are comparable to those higher for dolomite compared to limestone, especially at low tem­
observed in field treatments. There are no adjustable parameters in the perature. Therefore, a much larger acid volume is required for do­
model for scale-up from linear to radial flow, and ability of the model to lomites and the penalty from deviation from the optimum rate is also
give a reasonable prediction for radial flow strengthens the confidence much higher.
in the model. � For limestone the optimum PVBT increases monotonically with
Multiple field scale design examples are also presented. Key obser­ temperature, while for dolomites there is an optimum temperature as
vations from the study are: well as an optimum injection rate (Fig. 21 b).
� Surprisingly, emulsified acids are much more effective in reducing
the PVBT in dolomites than for limestones.

Appendix A

Calculation Example for HCl-Limestone System. In the following, we will show the required calculations to produce the simulated orange curve of

22
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Fig. 38. The change in acid optimum injection rate and acid volume with the wormhole length at 75, 150, and 300 � F. Emulsified acid-dolomite system.

Fig. 9b. The model parameters can be found in Table 2, for the HCl with corrosion inhibitor case. The experiments were extracted from Zakaria et al.,
(2015). The temperature was 150 � F, the cores dimensions were 1.5 in. diameter and 6 in. long. The cores were Indiana limestone with a porosity of
15%.
n ¼ 0.65.
a ¼ 5.10E-04 (m)2 (1–n).
b ¼ 35.1 (m/s)–1.
ko ¼ 2.43 Eþ06 m-2.
L ¼ 1 m.
T ¼ 60 s.
f ¼ 1.
Step 1: calculate the diffusion coefficient using Eq. (18), where CAo ¼ 0:15 ​ ​ & ​ ​ T ¼ 297:2 ​ K ​
� �
2270
Dm ¼ exp þ 1:326 � 0:15 12:11 ¼ 3:24E 09 m2=s
297:2

Step 2: calculate the dimensionless radius (r), where rc ¼ 1:5


2
� 0:0254 ¼ 0:01905 m
rc 0:01905 ​
r¼ ¼ ¼ 0:01905 m
L 1
Step 3: Calculate the injection face cross-sectional area (Ao )
Ao ¼ πrc2 ¼ π � 0:019052 ¼ 0:00114 ​ m2

Step 4: Calculate the Darcy velocity (vo ): (e.g qo ¼ 1:02E 08 m3 =s)


qo 1:02E 08 ​
vo ¼ ​ ¼ ¼ 8:943E 06 m=s
Ao 0:00114
Step 5: calculate the velocity in the wormhole (v) using Eq. (29)
vo 8:943E 06
v ¼ n ¼ ¼ 0:001547 ​ m=s
a AAo þ bvo 5:10E 04 � 0:001140:65
þ 35:1 � 8:943E
0:00114
06

Step 6: Calculate Damko


€hler number (Da), the dimensionless velocity (ω), and the dimensionless length (λ)
keff L k0 Dm L 2:43E þ 06 ​ � ​ 3:24E 09 � 1
Da ¼ ​ ¼ ¼ ¼ 5:0916
v v 0:001547

v 0:001547
ω¼ ¼ ¼ 172:99 ​
vo 8:943E 06

l
λ¼ ¼ 6 � 0:0254 ¼ ​ 0:1524
L
Step 7: Calculate the PVBT using Eq. (11).
ð1 ϕt Þ �
PVBT ¼ eDaλ 1
ϕt λCAo X ωDa

23
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

ð1 0:15Þ �
PVBT ¼ e5:0916�0:1524 1 ¼ ​ 0:6037
0:15 � 0:1524 � 0:15 � 0:5417 � 172:99 � 5:0916
Step 8: Repeat Steps 4 to 7 using different injection rates to generate the PVBT curve.
Step 9: Calculate the optimum injection rate (qopt ) using Eq. (34)
aq ¼ aAn keff lc

aq ¼ 5:10E 04x0:001140:65 x2:43E þ 06 ​ x ​ 3:24E 09x0:1524

aq ¼ 7:476E 09 m3 =s

bq ¼ eð 1 bkeff lc Þ
¼ eð 1 35:1�2:43Eþ06 ​ � ​ 3:24E 09�0:1524Þ
¼ ​ 0:353

1 þ 1:3 ​ bq 2:6 ​ b2q


qopt ¼ aq ¼ 3:46E 08 ​ m3 =s
1 þ 2:7 ​ bq

Appendix B. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.petrol.2019.106776.

References Fogler, H.S., Lund, K., McCune, C.C., et al., 1973. Dissolution of selected minerals in mud
acid. In: Presented at AIChE 74th National Meeting. New Orleans, Louisiana, USA,
11–15 March.
Aidagulov, G., Qiu, X., Brady, D., Abbad, M., et al., 2018. New insights into carbonate
Fredd, C.N., Fogler, H.S., 1996. Alternative stimulation fluids and their impact on
matrix stimulation from high-resolution 3D images of wormholes obtained in radial
carbonate acidizing. In: Presented at the SPE Formation Damage Control
acidizing experiments. In: Presented at the SPE Kingdom of Saudi Arabia Annual
Symposium. https://doi.org/10.2118/31074-MS. Lafayyette, Louisiana, USA, 14-15
Technical Symposium and Exhibition. https://doi.org/10.2118/192366-MS.
February. SPE-31074-MS.
Dammam, Saudi Arabia, 23–26 April. SPE-192366-MS.
Fredd, C.N., Fogler, H.S., 1998. Influence of transport and reaction on wormhole
Al-Ghamdi, A.H., Hill, D., Nasr-El-Din, H.A., et al., 2011. Acid diversion using
formation in porous media. AIChE J. 44 (9), 1933. https://doi.org/10.1002/
viscoelastic surfactants: the effects of flow rate and initial permeability contrast. In:
aic.690440902.
Presented at the SPE Annual Technical Conference and Exhibition. https://doi.org/
Fredd, C.N., Fogler, H.S., 1999. Optimum conditions for wormhole formation in
10.2118/142564-MS. Denver, Colorado, USA, 30 October–2 November. SPE-
carbonate porous media: influence of transport and reaction. SPE J. 4 (3), 196–205.
142564-MS.
https://doi.org/10.2118/56995-PA. SPE-56995-PA.
Ali, M.T., Nasr-El-Din, H.A., 2019. A robust model to simulate dolomite-matrix acidizing.
Furui, K., Burton, R., Burkhead, D., et al., 2012. A comprehensive model of high-rate
SPE Prod. Oper. 34 (1), 109–129. https://doi.org/10.2118/191136-PA. SPE-
matrix-acid stimulation for long horizontal wells in carbonate reservoirs: Part I–
191136-PA.
Scaling up core-level acid wormholing to field treatments. SPE J. 17 (1), 271–279.
Arashi, N., Hishinuma, Y., Narato, K., et al., 1982. Mass transfer to a parallel-plate
https://doi.org/10.2118/134265-PA.
catalyst. Int. Chem. Eng. 22 (3), 489C.
Gong, M., El-Rabba, A.M., 1999. Quantitative model of wormholing process in carbonate
Aris, Rutherford, 1999. Elementary Chemical Reactor Analysis. Dover Publications, Inc.,
acidizing. In: Presented at the Mid-continent Operations Symposium. https://doi.
0-486-40928.
org/10.2118/52165-MS. Oklahoma City, Oklahoma, USA, 18–21 February. SPE-
Bazin, B., 2001. From matrix acidizing to acid fracturing: a laboratory evaluation of acid/
52165-MS.
rock interactions. SPE Prod. Facil. 16 (1), 22–29. https://doi.org/10.2118/66566-
Hoefner, M.L., Fogler, H.S., 1989. Fluid-velocity and reaction-rate effects during
PA. SPE-66566-PA.
carbonate acidizing: application of network model. SPE Prod. Eng. 4 (1), 56–62. SPE-
Buijse, M.A., Glasbergen, G., 2005. A semi-empirical model to calculate wormhole
15573-PA. https://doi:10.2118/15573-PA.
growth in carbonate acidizing. In: Presented at the SPE Annual Technical Conference
Hung M., K., Hill D., A., Sepehrnoori, K., 1989. A mechanistic model of wormhole growth
and Exhibition. https://doi.org/10.2118/96892-MS. Dallas, Texas, USA, 9–10
in carbonate matrix acidizing and acid fracturing. JPT 41 (1), 59–66. https://doi.
October. SPE 96892-MS.
org/10.2118/16886-PA. In this issue.
Burton, R.C., Nozaki, M., Zwarich, N.R., et al., 2018. Improved understanding of acid
Kotb, A., Ali, M., Aziz Ezzat, A., et al., 2018. A computational fluid dynamics model for
wormholing in carbonate reservoirs through laboratory experiments and field
simulating the rotating disk apparatus. In: Presented at the SPE International Heavy
measurements. In: Presented at the SPE Annual Technical Conference and
Oil Conference and Exhibition. https://doi.org/10.2118/193739-MS. Kuwait City,
Exhibition. https://doi.org/10.2118/191625-MS. Dallas, Texas, USA, 24–26
Kuwait, 10–12 December. SPE 193739-MS.
September. SPE-191625-MS.
Livescu, S., Vissotski, A., Chaudhary, S., 2018. Improved acid placement modeling for
Cheng, H., Zhu, D., Hill, A.D., 2017. The effect of evolved CO2 on wormhole propagation
matrix acidizing optimization. In: Presented at Abu Dhabi International Petroleum
in carbonate acidizing. SPE Prod. Oper. 32 (3), 325–332. SPE-178962-PA. https://do
Exhibition & Conference. https://doi.org/10.2118/193119-MS. Abu Dhabi, UAE,
i:10.2118/178962-PA.
12–15 November. SPE-193119-MS.
Conway, M.W., Asadi, M., Penny, G.S., et al., 1999. A comparative study of straight/
Lund, K., Fogler, S., McCune, C., 1973. Acidization I. The dissolution of dolomite in
gelled/emulsified hydrochloric acid diffusivity coefficient using diaphragm cell and
hydrochloric acid. Chem. Eng. Sci. 28, 691–700. https://doi.org/10.1016/0009-
rotating disk. In: Presented at the SPE Annual Technical Conference and Exhibition.
2509(77)80003-1.
https://doi.org/10.2118/56532-MS. Houston, Texas, 3–6 October. SPE 56532-MS.
Lund, K., Fogler, S., McCune, C., et al., 1975. Acidization II. The dissolution of calcite in
Daccord, G., Touboul, E., Lenormand, R., 1989. Carbonate acidizing: toward a
hydrochloric acid. Chem. Eng. Sci. 28, 691–700.
quantitative model of the wormholing phenomenon. SPE Prod. Eng. 4 (1), 63–68.
Maheshwari, P., Balakotaiah, V., 2013. 3-D simulation of carbonate acidization with HCl:
https://doi.org/10.2118/16887-PA. SPE-16887-PA.
comparison with experiments. In: Presented at the SPE Production and Operations
Daccord, G., Lenormand, R., Li� etard, O., 1993. Chemical dissolution of a porous medium
Symposium. https://doi.org/10.2118/164517-MS. Oklahoma City, Oklahoma. SPE
by a reactive fluid—I. Model for the “wormholing” phenomenon. Chem. Eng. Sci. 48
164517-MS.
(1), 169–178. https://doi.org/10.1016/0009-2509(93)80293-Y.
McDuff, D., Jackson, S., Shuchart, C., et al., 2010. Understanding wormholes in
Dong, K., 2018. A new wormhole propagation model at optimal conditions for carbonate
carbonates: unprecedented experimental scale and 3D visualization. J. Pet. Technol.
acidizing. J. Pet. Sci. Eng. 171, 1309–1317. https://doi.org/10.1016/j.
62 (10), 78–81. https://doi.org/10.2118/129329-JPT. SPE-129329-JPT.
petrol.2018.08.055.
Panga, M.K.R., Balakotaiah, V., Ziauddin, M., 2002. Modeling, simulation and
Dong, K., Jin, X., Zhu, D., et al., 2014. The effect of core dimensions on the optimal acid
comparison of models for wormhole formation during matrix stimulation of
flux in carbonate acidizing. In: Presented at the SPE International Symposium and
carbonates. In: Presented at the SPE Annual Technical Conference and Exhibition.
Exhibition on Formation Damage Control. https://doi.org/10.2118/168146-MS.
https://doi.org/10.2118/77369-MS. San Antonio, Texas, 29 September–2 October.
Lafayette, Louisiana, USA, 26–28 February. SPE-168146-MS.
SPE 77369-MS.
Dong, K., Zhu, D., Hill, A.D., 2018. The role of temperature on optimal conditions in
Panga, M.K.R., Ziauddin, M., Balakotaiah, V., 2005. Two-scale continuum model for
dolomite acidizing: an experimental study and its applications. J. Pet. Sci. Eng. 165,
simulation of wormholes in carbonate acidization. AIChE J. 51 (12), 3231–3248.
736–742. https://doi.org/10.1016/j.petrol.2018.03.018.
https://doi.org/10.1002/aic.10574.
Economides, M.J., Hill, A.D., Ehlig-Economides, C., 1994. Petroleum Production
Systems, first ed. Prentice Hall, Upper Saddle River, NJ.

24
M. Ali and M. Ziauddin Journal of Petroleum Science and Engineering 186 (2020) 106776

Qiu, X.W., Zhao, W., Dyer, S.J., Al Dossary, A., et al., 2014. Revisiting reaction kinetics International Conference and Exhibition on Formation Damage Control. https://doi.
and wormholing phenomena during carbonate acidizing. In: Presented at the org/10.2118/189506-MS. Lafayette, Louisiana, USA, 7–9 February. SPE-189506-MS.
International Petroleum Technology Conference. https://doi.org/10.2523/IPTC- Tardy, P., Lecerf, B., Christanti, Y., 2007. An experimentally validated wormhole model
17285-MS. Doha, Qatar, 19-22 January. IPTC-17285-MS. for self-diverting and conventional acids in carbonate rocks under radial flow
Qiu, X., Edelman, E., Aidagulov, G., Ghommem, M., et al., 2018. Experimental conditions. In: Presented at the European Formation Damage Conference. https://
investigation of radial and linear acid injection into carbonates for well stimulation doi.org/10.2118/107854-MS. Scheveningen, The Netherlands, 30 May–1 June. SPE-
operations. In: Presented at the SPE Kingdom of Saudi Arabia Annual Technical 107854-MS.
Symposium and Exhibition. https://doi.org/10.2118/192261-MS. Dammam, Saudi Taylor, K.C., Nasr-El-Din, H.A., Mehta, S., 2006. Anomalous acid reaction rates in
Arabia, 23–26 April. SPE-192261-MS. carbonate reservoir rocks. SPE J. 11 (4), 488–496. https://doi.org/10.2118/89417-
Rabie, A.I., Nasr-El-Din, H.A., 2015. Effect of acid additives on the reaction of PA. SPE-89417-PA.
stimulating fluids during acidizing treatments. In: Presented at the SPE North Africa Wang, Y., Hill, D., Schechter, S., 1993. The optimum injection rate for matrix acidizing of
Technical Conference and Exhibition. https://doi.org/10.2118/175827-MS. Cairo, carbonate formations. In: Presented at the SPE Annual Technical Conference and
Egypt, 14–16 Septmber. SPE-175827-MS. Exhibition. https://doi.org/10.2118/26578-MS. Houston, Texas, USA, 3–6 October.
Roberts, L.D., Guin, J.A., 1975. A new method for predicting acid penetration distance. SPE-26578-MS.
SPE J. 15 (4), 277–286. https://doi.org/10.2118/5155-PA. SPE-5155-PA. Zakaria, A.S., Sayed, M., Nasr-El-Din, H.A., 2014. New insights into propagation of
Sayed, M.A.I., Nasr-El-Din, H.A., 2013. Acid treatments in high temperature dolomitic emulsified acids in vuggy dolomitic rocks. SPE J. 19 (1), 150–160. https://doi.org/
carbonate reservoirs using emulsified acids: a coreflood study. In: Presented at the 10.2118/163288-PA. SPE-163288-PA.
SPE Production and Operations Symposium. https://doi.org/10.2118/164487-MS. Zakaria, A., Nasr-El-Din, H.A., Ziauddin, M., 2015a. Flow of emulsified acid in carbonate
Oklahoma City, Oklahoma, USA, 23–26 March. SPE-164487-MS. rocks. Ind. Eng. Chem. Res. 54 (16), 4190–4202. https://doi.org/10.1021/
Sayed, M.A.I., Nasr-El-Din, H.A., Zhang, L., et al., 2012. A new emulsified acid to ie504167y.
stimulate deep wells in carbonate reservoirs: coreflood and acid reaction studies. In: Zakaria, A.S., Nasr-El-Din, H.A., Ziauddin, M., 2015b. Predicting the performance of the
Presented at North Africa Technical Conference and Exhibition. https://doi.org/ acid-stimulation treatments in carbonate reservoirs with nondestructive tracer tests.
10.2118/151062-MS. Cairo, Egypt, 20–22 February. SPE-151062-MS. SPE J. 20 (6), 1238–1253. https://doi.org/10.2118/174084-PA. SPE-174084-PA.
Seagraves, A.N., Smart, M.E., Ziauddin, M.E., 2018. Fundamental wormhole
characteristics in acid stimulation of perforated carbonates. In: Presented at the SPE

25

You might also like