You are on page 1of 9

SPE 116601

Matrix Acidizing of Carbonate Reservoirs Using Organic Acids and Mixture


of HCl and Organic Acids
F.F. Chang, SPE, Schlumberger; H.A. Nasr-El-Din, SPE, Texas A&M University; and T. Lindvig and X.W. Qiu,
Schlumberger


Copyright 2008, Society of Petroleum Engineers

This paper was prepared for presentation at the 2008 SPE Annual Technical Conference and Exhibition held in Denver, Colorado, USA, 2124 September 2008.

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


Abstract
Hydrochloric acid is the most commonly used acid for carbonate acidizing due to its low cost and high dissolving power.
However, there are two major drawbacks associated with using concentrated HCl solutions in deep wells. The first is its high
reaction rate with carbonate rocks, which limits acid penetration in the formation. The second is its corrosivity to well
tubulars. Hence organic acids become viable material for matrix acidizing to alleviate these two problems. Though organic
acids provide the benefit of retardation and low corrosivity, their low dissolving capacity may still limit the wormhole
penetration leading to insufficient stimulation of the formation. Therefore, opportunity exists to mix HCl with an organic
acid to achieve productivity enhancement by optimizing the wormhole penetration and profile.
Organic acids that are utilized in stimulating carbonate formations include formic, acetic, and more recently, citric and
lactic. Selecting a suitable organic acid for a specific acidizing treatment is more difficult due to complex thermodynamic
equilibrium and reaction kinetics. The reactions between organic acids and carbonate are less understood than those of HCl
with carbonate rocks. Organic acid/carbonate systems are complicated because of the presence of CO
2
, organic ligands, and
potential precipitation of the reaction products; the organic salts of calcium and magnesium. Therefore, more testing and
modeling are needed to better understand these reactions.
This paper discusses the required information to properly design an organic acid or HCl plus organic acid treatment. In
additional to reaction kinetics, data such as carbonate dissolving capacity at reservoir temperature and pressure, solubility of
reaction products, and the effect of HCl to organic acid ratio are needed to better design field treatments. Recommendations
are given on what and how laboratory evaluation should be carried out to obtain this information.

Introduction
Oil and gas companies are developing carbonate reservoirs of deeper and deeper depths in order to meet the demand of
increasing worldwide energy consumption. Enhancing productivity from these reservoirs poses a challenge in stimulation
fluids due to the increase in bottom hole temperature. The rapid reaction rate between HCl and carbonate limits the
penetration of HCl into the formation, especially at low pumping rates. The reaction of HCl often needs to be retarded by
gelling,
1
emulsifying,
2
or adding viscoelastic surfactants.
3
In addition to the high reaction rate, HCl is very corrosive to well
tubulars. Expensive corrosion inhibitors can protect the tubulars at high temperatures only for a short period of time. These
drawbacks make organic acids, such as formic and acetic, potentially attractive for stimulating high temperature wells.
Organic acids have been used in well stimulation because of their low corrosivity
4
and lower reaction rate with the rock.
However, they have the following limitations: (1) they cannot be used at high acid concentrations. This is because of the
limited solubility of their calcium salts. For instance, acetic and formic acids are typically used at concentrations less than 13
and 9 wt%, respectively to avoid precipitation of calcium acetate and calcium formate,
5
(2) Organic acids have a low
dissociation constant. They normally do not react to their full acid capacity because of the release of CO
2
from carbonate
dissolution, (3) the degree of hydrogen ion generation decreases with increasing temperature,
6,7
and (4) the cost of organic
acid is significantly higher than that of HCl for equivalent mass of rock dissolved.
Mixing HCl and an organic acid can provide several benefits than using either acid alone.
8-10
Organic acids, such as
formic acid, can be used as a corrosion inhibitor intensifier for high temperature applications. More interestingly, organic
acids can be used to enhance acid penetration in the formation. When an organic acid is mixed with HCl, it does not
2 SPE 116601
dissociate to generate hydrogen ions due to its low dissociation constant. Therefore, the organic acid is preserved until HCl is
nearly spent. As the hydrogen ions from HCl are depleted, the carboxylic groups start to dissociate, resulting in further
carbonate rock dissolution from the tip of the acid front and hence increasing acid penetration into the formation. However,
there have been claims that when dissolving carbonate with an HCl-organic acid system, the conversion of organic acid is
further reduced. For example, in a modeling study by Buijse et al.
11
it was illustrated that in a 15 wt% HCl/10 wt% acetic
acid system, only 24% of the 10 wt% acetic acid was spent. This result was believed to be caused by the large amounts of
CO
2
generated from the reaction of HCl with carbonates, and the reversibility of organic (HA) acid reacting with carbonates,
Eq. 1.
7


CaCO
3
+ 2 HA CaA
2
+ H
2
O + CO
2
(1)

In Eq. 1, increasing CO
2
concentration drives the reaction toward the left, therefore less CaCO
3
is dissolved. There is very
limited experimental data to prove the validity of the models described by William et al.,
7
and Buijse et al.
11

Carbon dioxide also affects the reactions of organic acids via the formation of carbonic acid. Under reservoir pressure,
the CO
2
in most cases is dissolved in the aqueous phase and generates carbonic acid, Eqs. 2-4. Carbonic acid buffers the pH
value to nearly 4.5:

H
2
O + CO
2
H
2
CO
3
(2)

H
2
CO
3
H
+
+ HCO
3
-
(3)

HCO
3
-
H
+
+ CO
3
2-
(4)


Figs. 1(a) and 1(b) show the formic acid and acetic acid dissociation as a function of pH. It can be seen that at pH 4.5,
formic and acetic acids are not fully dissociated.
Acid-carbonate reaction kinetics has been studied by several research groups.
12-16
The reaction rate and mass transfer
coefficients between various acids and calcite and dolomite are well documented. However, the equilibrium chemistry of the
acid-carbonate rock system is thought to be simple, therefore it has not been emphasized. For HCl, the acid spending is to
completion. Therefore, its total rock dissolving capacity is straight forward. It can be calculated stoichiometrically by
rewriting Eq. 1 as:

CaCO
3
+ 2 HCl CaCl
2
+ H
2
O + CO
2
(5)

The acid reaction with calcite can be written as:

CaCO
3
+ 2 H
+
Ca
2+
+ H
2
O + CO
2
(6)

Unlike HCl, organic acids and limestone reaction reaches equilibrium before the acid is completely spent. The total acid
dissolving capacity requires an accurate account of the thermodynamic parameters of a system of reactions including CO
2

and carbonic acid equilibrium. The system can be further complicated by the ion association of calcium ions and the anions
of the organic acid, and potential precipitation of the calcium salt with the anion, A
-
, of the organic acid, Eqs. 7 and 8:

Ca
2+
+ A
-
CaA
+
(7)
Ca
2+
+ 2A
-
CaA
2
(8)

The equilibrium constants of these reactions, especially at reservoir temperature and pressure, are not available for all
reactions in the system. Therefore, it is difficult to properly model spending of organic acids. A model based on Gibbs free
energy is required to study the thermodynamic equilibrium among the species in solution.
The objective of this paper is to investigate the equilibrium of the system of reactions involved in the limestone
dissolution by organic acids and HCl-organic acid solutions. The effect of reservoir pressure on CO
2
evolved due to calcite
dissolution by acid is investigated. Commercial chemical analysis software is used to establish the equilibrium of ionic
species in the reactions under reservoir temperature and pressure. The software has been applied for chemical analyses in
various industries. It has extensive database of thermodynamic parameters and been widely validated for its accuracy and
functionality against a large number of experimental data sets. This approach can help select proper experimental parameters
and design acid formulation for carbonate acidizing treatments. It complements the kinetic studies of acid-rock reaction rate.


SPE 116601 3
Equilibrium Modeling
In an acid-carbonate system, the electrolyte concentrations are at high levels, and the solution cannot be treated as an idea
one. An ideal solution means that the concentration approaches zero or the mole fraction approaches unity.
17
Therefore,
when calculating thermodynamic equilibrium, the activity of each species has to be used. The chemical analysis model
utilized in this study uses the Helgeson model
18
to represent the variation of species standard Gibbs free energies with
temperature and pressure and a Pitzer-like activity coefficient model
19
to account for deviations from thermodynamic
ideality. The advantage of a Gibbs energy free energy equilibrium framework is that reaction constants are not required. The
software considers all the relevant species and manifests their equilibrium according to their free energy functions.
Four acid formulations were tested in the simulation to investigate their interactions with calcite rock. These acid
formulations were: (1) 10 wt% acetic acid, (2) 15 wt% HCl + 10 wt% acetic acid, (3) 9 wt% formic acid, and (4) 15 wt% HCl
+ 9 wt% formic acid. The simulation considered a reaction vessel that contained 100 kg of the acid, calcite was sequentially
added into the vessel at 1 kg increment to react with the acid at 150
o
F and 1,000 psi. The distribution of each chemical
species was calculated at each step based on chemical equilibrium.
Fig. 2 shows the case of 10 wt% acetic acid reacting with calcite. As CH
3
COOH reacted with CaCO
3
, the CO
2

concentration in the aqueous increased and Ca
2+
was released. After 6.2 kg of CaCO
3
was dissolved, the system reached
equilibrium. At the end of the reaction, 74% of the 10 wt% acetic acid was spent. This number was high compared to the
54% obtained from the model predictions by Buijse et al.
11
and the 50% obtained from the experimental results of Chatelain
et al.
6
At 1,000 psi system pressure and 150
o
F temperature, the initial pH of 10 wt% acetic acid is 2.32. The pH at the end of
the reaction was 4.48 due to dissolved CO
2
in the system. It was notable that in the system the free Ca
2+
concentration was
significantly lower than the stoichiometry according to reaction (6). And the concentrations of acetate ions (CH
3
COO
-
),
calcium acetate (Ca(CH
3
COO)
2
), and calcium monoacetate (Ca(CH
3
COO)
+
) were in the same order of magnitude and not
negligible. This indicated that Ca
2+
ions and the acetate ions were in ionic association and the reactions (7) and (8) could not
be ignored.
20,21
This was the reason that the model predicted higher dissociation of acetic acid than previous publication by
Buijse et al.
11
and Chatelain et al.
6
However, Bombardieri and Martin
22
showed that at 150
o
F and 300 psi 65% of the 10 wt%
acetic acid reacted with calcium carbonate in 7 hours while the dissolution was still progressing. Though the test by
Bombardieri and Martin
22
was at lower pressure level, data showed that at 300 psi, the CO
2
dissolved in aqueous solution was
sufficient to buffer the pH to 4.3, which was not much different from the level at 1,000 psi. Hence the reduction of acid
spending due to CO
2
should be limited. On the other hand, increasing organic acid spending due to Ca
2+
may be significant.
When reactions (7) and (8) were dominant, the organic acid dissociation could be further driven to the right to increase the
amount of hydrogen ion generation as follows:

HA H
+
+ A
-
(9)

When the free organic anion (A
-
) was reduced due to its association with Ca
2+
by reactions (7) and (8), the equilibrium of
reaction (9) shifted toward the right and more H
+
was produced. This was demonstrated by a simple experiment of pH
measurement. The pH of a 100 cm
3
solution of 10 wt% acetic acid was first measured at ambient condition. The pH of the
solution was continuously monitored while CaCl
2
was slowly added to the solution. Fig. 3 shows the pH of the 10 wt%
acetic acid solution decreased as more CaCl
2
was added to the solution, indicating the acetic acid further dissociated and
more H
+
was generated. Buijses model
11
did not account for the effect of ionic association between Ca
2+
and A
-
therefore it
may be underestimating spending of the organic acid. The data presented by Chatelain et al.
6
was obtained from
experiments. It is difficult to argue against the validity of their results without conducting additional experiments.
Nonetheless, it is worth noting that the experimental procedure used by Chatelain et al.
6
to measure acid concentration was by
first flushing the acid through the carbonate rock, then reducing the pressure of the acid effluent to allow CO
2
to escape. The
effluent was then titrated with NaOH solutions. There was still dissolved CO
2
in the solution so that carbonic acid
contributed to the residual acidity. No material balance was conducted to compare the acid spending with rock weight loss,
nor pH measurement to demonstrate the complete vaporization of CO
2
. Therefore, it is the belief of the authors that more
experimental work is required to properly define the capacity of organic acids under downhole conditions.
The next simulation was related to calcite dissolution by a mixture of 15 wt% HCl and 10 wt% acetic acid. It is well
known that when an organic acid is mixed with HCl, the organic acid will not dissociate due to high hydrogen ion
concentration provided by HCl. It is when HCl is nearly completely spent that the organic acid starts to dissociate. This is
beneficial to the acid penetration depth because the strength of the organic acid is preserved. It behaves as if the fresh
organic acid were injected at the tip of HCl reaction front inside the formation. Another benefit of the mixed acid system is
that the reaction rate is reduced.
23
However, there have been claims
7,11
that the organic acid spending percentage will be
further reduced when it is mixed with HCl due to large amounts of CO
2
generated by the HCl-CaCO
3
reaction.
Fig. 4 shows the chemical species distribution as the acid mixture dissolved calcite. The concentration of CH
3
COOH
remained relatively constant when fresh HCl was reacting with the rock. After the 15 wt% HCl was fully spent dissolving
about 20 kg of rock, the acetic acid started to react and the acid concentration declined sharply. Totally 28.1 kg of CaCO
3

was dissolved by the 100 kg of 15 wt% HCl-10 wt% acetic acid mixture. The chemical analysis shown in Fig. 4 indicates
that the effect of CO
2
was less pronounced, the pH at the end of reaction was 4.29, which was not much lower than the 10
wt% acetic acid case. However, the Ca
2+
plays a key role in the reaction. As the Ca
2+
increased due to HCl dissolving the
4 SPE 116601
calcite, a large amount of calcium monoacetate was formed at the expense of acetate ions (CH
3
COO
-
) and calcium acetate
(Ca(CH
3
COO)
2
). The CH
3
COO
-
and Ca(CH
3
COO)
2
concentrations were reduced significantly as shown in Fig. 4. This
means reaction (7) was predominant due to high Ca
2+
concentration, which droves the acetic acid dissociation toward the
right. At equilibrium, the acetic acid spending percentage reached 90%. There has not been published experimental data
concerning the equilibrium during carbonate rock dissolution by HCl and organic mixtures. Research is underway to validate
the simulation and to gain further understanding in this area.
Fig. 5 shows the case of 9 wt% formic acid reacting with calcite. Generally, the trend was similar to that of acetic acid
(Fig. 2). At 1,000 psi system pressure and 150
o
F, the initial pH of 9 wt% formic acid was 1.74. As the acid reacted with
calcite, the pH and dissolved CO
2
concentration increased. The CO
2
in solution reached its solubility limit of 2.86% at 1,000
psi when pH increased to 3.45 after 7 kg of calcite was dissolved. Formic acid continued to dissociate until 97% of the
formic acid was spent and a total of 9.7 kg of calcite was dissolved when the system reached equilibrium. An additional 1%
of CO
2
was generated in the form of gas. The pH at the end of the reaction was 4.48. The 97% spending was again higher
than 92% from the experimental result by Chatelain et al.
6
and 85% from the model calculation by Buijse et al.
11
It was
noticed that there was no free Ca
2+
in the solution. All the Ca
2+
were present in the form of calcium monoformate
Ca(HCOO)
+
and calcium formate Ca(HCOO)
2
. There was no precipitation as the two calcium salts were in the aqueous
phase. The concentrations of free calcium ions (Ca
2+
), formate ions (HCOO
-
), calcium formate (Ca(HCOO)
2
), and calcium
monoformate (Ca(HCOO)
+
) also explained that reactions (7) and (8) were prevalent.
The final case studied was calcite dissolution by 15 wt% HCl and 9 wt% formic acid mixture. Fig. 6 shows the chemical
species distribution as the acid mixture dissolved calcite. Similar to the HCl and acetic acid mixture, the concentration of
HCOOH remained relatively constant until HCl reacted to near completion when 20 kg of calcite was dissolved. The formic
acid concentration then started to decline sharply as it dissociated into formate and hydrogen ions to react with the rock. At
the end of the reaction, the total CaCO
3
dissolved was 30.4 kg by the 100 kg of 15 wt% HCl-9 wt% formic acid mixture. The
pH at the end of reaction was 4.33, slightly lower than that by 9 wt% formic acid alone. The large amount of free Ca
2+
ions
from HCl-calcite reaction was consumed by formate ions to form calcium monoformate by ionic association. Hence, the free
formate ions (HCOO
-
) and calcium formate (Ca(HCOO)
2
) was depleted from the system. Reaction (7) dominated and it
drove the formic acid dissociation toward the right. At equilibrium, the formic acid spending reached 99%.
It has been known that organic acid spending is significantly impacted by the CO
2
in solution. A common belief
6,11
in the
oil and gas industry is that at 1,000 psi the CO
2
generated by the acid-carbonate reaction will be dissolved in the solution.
Therefore 1,000 psi has been the standard system pressure used in experiments related to carbonate dissolution. A simulation
was conducted to illustrate the effect of system pressure on organic acid spending and CO
2
solubility. The acid formulation
for the simulation run was 15 wt% HCl and 9 wt% formic acid mixture. The acid reacted with calcite at 150
o
F and pressure
ranging from 1 to 6,000 psi. Fig. 7 shows the pH of the system at equilibrium and the total CO
2
in aqueous and vapor phases.
It can be seen that a large fraction of the CO
2
generated by the acid-calcite reaction was in the vapor phase below 3,000 psi.
The amount of dissolved CO
2
increased dramatically when the pressure was above 3,000 psi until all the vapor phase was
dissolved in the aqueous solution when the pressure was above 5,200 psi. This demonstrated that 1,000 psi was not sufficient
to keep all the produced CO
2
in solution. This has been pointed out by a few researchers in the past.
24,25
Interestingly
however, the equilibrium pH remained relatively constant as long as the pressure was above 200 psi and below 3,000 psi,
especially between 1,000 and 3,000 psi. This means the acid spending due to the CO
2
effect was small. Therefore, from acid
dissolving power aspect, application of 1,000 psi system pressure should produce reasonable results. However, when
conducting kinetics experiments, CO
2
vapor significantly affects the reaction rate due to the forced convection effect to
promote mixing and mass transfer in the reactive boundary layer. Therefore, the reaction rate of acid on carbonates can be
overestimated if insufficient system pressure is applied.
24


Conclusions
The chemical analysis software based on the Gibbs free energy is a very useful tool to gain insight into organic acid reactions
with carbonate rocks. The objective of this paper is to highlight that further research, especially in experimentation, is
needed to understand the role of organic acids in acidizing deep wells. Based on the work discussed in this paper, the
following conclusions can be drawn:

1. When conducting carbonate acidizing experiments, 1,000 psi system pressure may be sufficient for testing acid
dissolving capacity, but a much higher pressure is required for kinetics studies.
2. The ionic association between Ca
2+
and organic anions can significantly increase organic acid dissociation.
3. Mixing HCl with an organic acid (acetic or formic) provides benefits not only in reducing corrosivity, but also in
acid penetration depth due to delayed dissociation of organic acid in the presence of live HCl.




SPE 116601 5
References
1. Nasr-El-Din H.A., Al-Mohammad, A.M., Al-Aamri, A.M., and Al-Fuwaires, O.: Reaction Kinetics of Gelled Acids
with Calcite, paper SPE 103979, accepted for publication in SPEPO, 2008.
2. Navarrete, R.C., Holms, B.A., McConnell, S.B. and Linton, D.E.: Emulsified Acid Enhances Well Production in High-
Temperature Carbonate Formations, paper SPE 50612 presented at the 1998 SPE European Petroleum Conference, The
Hague, The Netherlands, October 20 - 22.
3. Al-Mohammad, A.M., Nasr-El-Din H.A., Al-Aamri, A.M., and Al-Fuwaires, O.: Reaction of Calcite with Surfactant-
Based Acids, paper SPE 102838 presented at the 2006 SPE Annual Technical Conference and Exhibition, San Antonio,
TX, 24-27 September.
4. Harris, F.N.: Applications of Acetic Acid to Well Completions, Stimulations and Reconditioning, JPT 13(7) (1961)
637-639.
5. Robert, J.A. and Crowe, C.W.: Carbonate Acidizing Design, Reservoir Stimulation, Economides, M.J. and Nolte, K.G.
3
rd
Edition, John Wiley & Sons Inc. (2000) 17-11.
6. Chatelain, J.C. Silberberg, I.H., and Schechter, R.S.: Thermodynamic Limitations in Organic-Acid/Carbonate
Systems, SPEJ 16(4) (1976) 189-195.
7. Williams, B.B., Gidley, J.L., and Schechter, R.S.: Acidizing Fundamentals, SPE Monograph Volume 6, 1979.
8. Dill, W.R. and Keeney, B.R.: Optimizing HCl-Formic Acid Mixtures for High Temperature Stimulation, paper SPE
7567 presented at the 1978 Annual Fall Technical Conference and Exhibition of the SPE-AIME, Houston, TX, Oct. 1-3.
9. Taylor, K.C., Al-Katheeri, M.I., and Nasr-El-Din, H.A.: Development and Field Application of a New Measurement
Technique for Organic Acid Additives in Stimulation Fluids, SPEJ 10(2) (2005) 152-160.
10. Katheeri, M.I., Nasr-El-Din, H.A., Taylor, K.C., and Grainees, A.H.: Determination and Fate of Formic Acid in High
Temperature Acid Stimulation Fluids, paper SPE 73749 presented at the 2002 SPE International Symposium on
Formation Damage Control, Lafayette, LA, Feb. 20-21.
11. Buijse, M., deBoer, P., Breukel, B., Klos, M., and Burgos, G.: Organic Acids in Carbonate Acidizing, SPEPF 19(3)
(2004) 128-134.
12. Fredd, C.N. and Fogler, H.S.: The Kinetics of Calcite Dissolution in Acetic Acid Solution, Chem. Eng. Sci. 53(22)
(1998) 3863-3874.
13. de Rozieres, J., Chang, F.F. and Sullivan, R.B.: Measuring Diffusion Coefficients in Acid Fracturing Fluids and Their
Application to Gelled and Emulsified Acids, paper SPE 28552 presented at the 1994 Annual SPE Conference, New
Orleans, LA, Sep. 25-28.
14. Nierode, D.E. and Williams, B.B.: Characteristics of Acid Reaction in Limestone Formations, SPEJ 11 (1971) 406-
418.
15. Roberts, L.D. and Guin, J.A.: The Effect of Surface Kinetics in Fracture Acidizing, SPEJ 14 (1974) 385-395.
16. Nasr-El-Din, H.A., Al-Mohammad, A.M., Al-Aamri, A.D., Al-Fahad, M.A. and Chang, F.F.: Quantitative Analysis of
Reaction Rate Retardation in Surfactant-Based Acids, paper SPE 107451, accepted for publication in SPEPO, 2008.
17. Laitinen, H.A. and Harris, W.E.: Chemical Analysis, 2
nd
edition, 1975, McGraw-Hill Book Company.
18. Shock, E.L., Helgeson, H.C., and Sverjensky, D.A.: Calculation of the Thermodynamic Properties of Aqueous Species
at High Pressures and Temperatures: Standard Partial Molal Properties of Inorganic Neutral Species, Geochim.
Cosmochim. Acta 53 (1989) 2157-2183.
19. Pitzer, K.S.: Thermodynamics of Electrolytes. I. Theoretical Basis and General Equation, J. Phys. Chem. 77 (1973)
268-277.
20. Loos, D., Pasel, C., Luckas, M., Schmidt, K.G., and Herbell, J.D.: Experimental Investigation and Modelling of the
Solubility of Calcite and Gypsum in Aqueous Systems at Higher Ionic Strength, Fluid Phase Equilibrium 219 (2004)
219-229.
21. Nancollas, G.H.: Thermodynamics of Ion Association, Part II Alkaline-Earth Acetates and Formates, J. Chem. Soc.
(1956) 744.
22. Bombardieri, C.C. and Martin, T.H.: Acid Treating Process, US Patent 3,251,415, May 17, 1966.
23. Dill, W.R.: Reaction Time of Hydrochloric-Acetic Acid Solutions on Limestone, paper SPE 00211 presented at the
16th Southwest Regional Meeting of the American Chemical Society, Oklahoma City, OK, December 1-3, 1960.
24. Mumallah, N.A.: Factors Influencing the Reaction Rate of Hydrochloric Acid and Carbonate Rock, paper SPE 21036
presented at the 1991 SPE International Symposium on Oilfield Chemistry, Anaheim, CA, Feb. 20-22.
25. Crowe, C.W., McGowan, G.R., and Baranet, S.E.: Investigation of Retarded Acids Provides Better Understanding of
Their Effectiveness and Potential Benefits, SPEPE 5(2) (1990) 166-170.

6 SPE 116601
0.00
0.20
0.40
0.60
0.80
1.00
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
pH
F
r
a
c
t
i
o
n

o
f

S
p
e
c
i
e
s
HCOOH
HCOO
-

Fig. 1a: The fraction of species from the dissociation of formic acid at ambient
conditions.
0.00
0.20
0.40
0.60
0.80
1.00
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0
pH
F
r
a
c
t
i
o
n

o
f

S
p
e
c
i
e
s
CH
3
COOH CH
3
COO
-

Fig. 1b: The fraction of species from the dissociation of acetic acid at ambient conditions.
SPE 116601 7
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 1 2 3 4 5 6 7 8 9 10
Calcite Dissolved by 100 kg of Acid (kg)
C
o
n
c
e
n
t
r
a
t
i
o
n

(
m
o
l
/
l
)
0
1
2
3
4
5
p
H
CH3COOH
Ca2+
CH3COO-
Ca(CH3COO)+
Ca(CH3COO)2
CO2
pH

Fig. 2: Equilibrium concentrations of main species when 10 wt% acetic acid reacted with calcite at 150
o
F and 1,000 psi.

1
1.2
1.4
1.6
1.8
2
2.2
0 5 10 15 20 25 30
Weight of CaCl
2
Added, g
p
H

Fig. 3: Effect of calcium ion concentration on the pH of 100 cm
3
of 10 wt% acetic acid at ambient conditions.


8 SPE 116601
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 5 10 15 20 25 30 35
Calcite Dissolved by 100 kg of Acid (kg)
C
o
n
c
e
n
t
r
a
t
i
o
n

(
m
o
l
/
l
)
-1
0
1
2
3
4
5
p
H
CH3COOH
Ca2+
CH3COO-
Ca(HCOO)+
Ca(CH3COO)2
CO2
pH

Fig. 4: Equilibrium concentrations of main species when 15 wt% HCl and 10 wt% acetic acid reacted with calcite at 150
o
F
and 1,000 psi.
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 2 4 6 8 10 12
Calcite Dissolved by 100 kg of Acid (kg)
C
o
n
c
e
n
t
r
a
t
i
o
n

(
m
o
l
/
l
)
0
1
2
3
4
5
p
H
HCOOH
Ca2+
HCOO-
Ca(HCOO)+
Ca(HCOO)2
CO2
pH

Fig. 5: Equilibrium concentrations of main species when 9 wt% formic acid reacted with calcite at 150
o
F and 1,000 psi.
SPE 116601 9
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 5 10 15 20 25 30 35
Calcite Dissolved by 100 kg of Acid (kg)
C
o
n
c
e
n
t
r
a
t
i
o
n

(
m
o
l
/
l
)
-1
0
1
2
3
4
5
p
H
HCOOH
Ca2+
HCOO-
Ca(HCOO)+
Ca(HCOO)2
CO2
pH

Fig. 6: Equilibrium concentrations of main species when 15 wt% HCl and 9 wt% formic acid reacted with calcite at 150
o
F
and 1,000 psi.
0
1
2
3
4
5
0 1000 2000 3000 4000 5000 6000
Pressure (psi)
C
O
2

(
a
q
)

(
m
o
l
/
l
)

a
n
d

p
H
0
50
100
150
200
250
300
350
C
O
2

v
a
p
o
r

(
m
o
l
e
)
pH
CO2 (aq)
CO2 vapor


Fig. 7: pH value and CO
2
phase distribution as a function of pressure when 15 wt% HCl + 9 wt% formic acid reacted with
calcite at 150
o
F.

You might also like